首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   3710篇
  免费   312篇
  国内免费   5篇
  2023年   18篇
  2022年   43篇
  2021年   99篇
  2020年   50篇
  2019年   68篇
  2018年   94篇
  2017年   72篇
  2016年   140篇
  2015年   195篇
  2014年   223篇
  2013年   290篇
  2012年   300篇
  2011年   290篇
  2010年   183篇
  2009年   163篇
  2008年   204篇
  2007年   204篇
  2006年   200篇
  2005年   211篇
  2004年   194篇
  2003年   178篇
  2002年   164篇
  2001年   36篇
  2000年   25篇
  1999年   31篇
  1998年   52篇
  1997年   34篇
  1996年   25篇
  1995年   25篇
  1994年   14篇
  1993年   19篇
  1992年   19篇
  1991年   17篇
  1990年   11篇
  1989年   12篇
  1988年   14篇
  1987年   13篇
  1986年   5篇
  1985年   11篇
  1984年   10篇
  1983年   6篇
  1982年   5篇
  1981年   5篇
  1980年   13篇
  1979年   7篇
  1977年   4篇
  1974年   4篇
  1973年   3篇
  1965年   4篇
  1962年   3篇
排序方式: 共有4027条查询结果,搜索用时 562 毫秒
991.
Two mouse models of pelvic organ prolapse have been generated recently, both of which have null mutations in genes involved in elastic fiber synthesis and assembly (fibulin 5 and lysyl oxidase-like 1). Interestingly, although these mice exhibit elastinopathies early in life, pelvic organ prolapse does not develop until later in life. In this investigation we developed and validated a tool to quantify the severity of pelvic organ prolapse in mice, and we used this tool prospectively to study the role of fibulin 5, aging, and vaginal proteases in the development of pelvic organ prolapse. The results indicate that >90% of Fbln5(-/-) mice develop prolapse by 6 mo of age, even in the absence of vaginal delivery, and that increased vaginal protease activity precedes the development of prolapse.  相似文献   
992.
This study was conducted to shed light on the so far unexplored intracellular mechanisms underlying negative modulation of Leydig cell steroidogenesis by histamine (HA). Using the MA-10 cell line and highly purified rat Leydig cells as experimental models, we examined the effect of the amine on biochemical steps known to be modulated by HA or involved in LH/hCG action. In agreement with previous findings, HA at 10 microM showed a potent inhibitory effect on hCG-stimulated steroid synthesis, regardless of the gonadotropin concentration used. Moreover, HA decreased not only LH/hCG-induced cAMP production but also steroid synthesis stimulated by the permeable cAMP analog dibutyryl cAMP (db-cAMP). Considering the post-cAMP sites of HA action, it is shown herein that HA markedly inhibited db-cAMP-stimulated steroidogenic acute regulatory (STAR) protein expression, as well as steps catalyzed by P450-dependent enzymes, mainly the conversion of cholesterol to pregnenolone by cholesterol side-chain cleavage enzyme (CYP11A). The antisteroidogenic action of HA was blocked by addition of the phospholipase C (PLC) inhibitor U73122, and HA significantly augmented inositol triphosphate (IP3) production, suggesting a major role for the PLC/IP3 pathway in HA-induced inhibition of Leydig cell function. Finally, HA increased nitric oxide synthase (NOS) activity, and the NOS inhibitor NG-nitro-L-arginine methyl ester (L-NAME) markedly attenuated the effect of the amine on steroid synthesis. On the basis of our findings, HA antagonizes the gonadotropin action in Leydig cells at steps before and after cAMP formation. NOS activation is the main intracellular mechanism by which HA exerts its antisteroidogenic effects.  相似文献   
993.
Peroxiredoxin 2 (Prx2) is a 2-Cys peroxiredoxin extremely abundant in the erythrocyte. The peroxidase activity was studied in a steady-state approach yielding an apparent KM of 2.4 μM for human thioredoxin and a very low KM for H2O2 (?0.7 μM). Rate constants for the reaction of peroxidatic cysteine with the peroxide substrate, H2O2 or peroxynitrite, were determined by competition kinetics, k2 = 1.0 × 108 and 1.4 × 107 M−1 s−1 at 25 °C and pH 7.4, respectively. Excess of both oxidants inactivated the enzyme by overoxidation and also tyrosine nitration and dityrosine were observed with peroxynitrite treatment. Prx2 associates into decamers (5 homodimers) and we estimated a dissociation constant Kd < 10−23 M4 which confirms the enzyme exists as a decamer in vivo. Our kinetic results indicate Prx2 is a key antioxidant enzyme for the erythrocyte and reveal red blood cells as active oxidant scrubbers in the bloodstream.  相似文献   
994.
Cecilia Herraiz 《FEBS letters》2009,583(19):3269-3274
Melanocortin 1 receptor (MC1R), a Gs protein-coupled receptor expressed in melanocytes, is a major determinant of skin pigmentation, phototype and cancer risk. MC1R activates cAMP and mitogen-activated protein kinase ERK1/ERK2 signalling. When expressed in rat pheochromocytoma cell line cells, the R151C, R160W and D294H MC1R variants associated with melanoma and impaired cAMP signalling mediated ERK activation and ERK-dependent, agonist-induced neurite outgrowth comparable with wild-type. Dose-response curves for ERK activation and cAMP production indicated higher sensitivity of the ERK response. Thus, the melanoma-associated MC1R mutations impact differently on cAMP and ERK signalling, suggesting that cAMP is not responsible for functional coupling of MC1R to the ERK cascade.  相似文献   
995.
996.
In this report we investigate the alterations of the dielectric properties of the plasma membrane caused by the infection of cultured fibroblasts with murine polyomavirus. The approach consists in a well-established dielectric spectroscopy technique, electrorotation, which has been successfully used in our laboratory to study the alterations of the plasma membrane of cells exposed to various forms of stress. The response to viral proliferation was time dependent as shown by evaluation of the de novo synthesis of viral DNA. This response was paralleled by gradual damage of the membrane evidenced by alteration of the dielectric parameters, specific capacitance and conductance. The electrorotation results show a reduced effect on the dielectric properties of the membrane when infection is carried out in the presence of a natural oil (MEX). In this case a drastic reduction in viral DNA synthesis was also monitored, thus indicating an antiviral action of this product.  相似文献   
997.
Poribacteria were found in nine sponge species belonging to six orders of Porifera from three oceans. Phylogenetic analysis revealed four distinct poribacterial clades, which contained organisms obtained from several different geographic regions, indicating that the distribution of poribacteria is cosmopolitan. Members of divergent poribacterial clades were also found in the same sponge species in three different sponge genera.Recently, a novel bacterial phylum, termed “Poribacteria,” was discovered, and members of this phylum have been found exclusively in sponges (2). Phylogenetic analyses of 16S rRNA genes indicated that poribacteria are evolutionarily deeply branching organisms and related to a superphylum composed of Planctomycetes, Verrucomicrobia, and Chlamydia (11). Poribacterial 16S rRNA genes contain 13 of 15 planctomycete signature nucleotides, but a level of sequence divergence of more than 25% compared to any other bacterial phylum, including the Planctomycetes, justifies the status of this taxon as an independent phylum. A consistent treeing pattern is difficult to resolve in comparative phylogenetic sequence analyses, making the poribacteria an unusual line of phylogenetic descent. In addition to their divergent status as a separate phylum on the basis of the 16S rRNA sequence, poribacteria are also divergent because they may have a compartmentalized cell structure, a cell plan they share only with members of the phyla Planctomycetes and Verrucomicrobia (2). They are also of interest for understanding the potential contribution of obligate sponge-associated bacteria to the sponges harboring them and as an example of a yet-to-be-cultured group of bacteria associated with invertebrate tissue apparently exclusively but for unknown reasons. This study aimed to further explore the presence and diversity of poribacteria in different marine demosponge genera using samples from around the world.The Mediterranean sponges were collected by scuba divers offshore at Banyuls sur Mer, France (42°29′N, 03°08′E). The Caribbean sponges were collected offshore at Little San Salvador Island, Bahamas (24°32′N, 75°55′W). The eastern Pacific sponge Aplysina fistularis was collected offshore at San Diego, CA (32°51′N, 117°15′W). The western Pacific sponge Theonella swinhoei was collected offshore at Palau (07°23′N, 134°38′E). All non-Great Barrier Reef (non-GBR) sponges were collected between May and July 2000, and once individual sponge specimens were brought to the surface, they were frozen in liquid nitrogen on board ship and stored at −80°C until microbiological processing (9). The GBR marine sponges were collected off Heron Island Research Station (23°27′S, 151°5′E) in April 2002 (5). Pseudoceratina clavata was collected by scuba divers at a depth of 14 m, and Rhabdastrella globostellata was collected at a depth of ca. 0.5 m after a reef walk consisting of a few hundred meters. The samples were immediately placed in plastic bags and brought to Heron Island Research Station, where they were stored at −80°C until processing. Sponge DNA was extracted as described previously (2, 5).Total sponge-derived genomic DNA was screened by PCR for the presence of poribacteria using a 16S rRNA gene primer set. Poribacterial 16S rRNA genes were amplified by employing a pair of Poribacteria-specific primers, POR389f (5′-ACG ATG CGA CGC CGC GTG-3′) and POR1130r (5′-GGC TCG TCA CCA GCG GTC-3′) (2). The poribacterial PCR products that were ca. 740 bp long derived from one sponge individual were cloned into the pGEM-T Easy vector (Promega, Madison, WI). Clone inserts were digested with restriction endonucleases MspI and HaeIII (New England Biolabs, Inc., United States), characterized to obtain restriction profiles and unique profiles, and sequenced. The compiled partial 16S rRNA gene sequences were then analyzed using BLASTN to select the most closely related poribacterial reference sequences.The sequences exhibiting levels of similarity of less than 97% were used for further analysis. Poribacterial 16S rRNA gene sequences were aligned using the ARB software package (7). The resulting alignment was imported into PAUP (10) and analyzed by using distance, maximum parsimony, and maximum likelihood algorithms together with bootstrap resamplings (3,000, 3,000, and 200 resamplings, respectively), and the resulting bootstrap values were applied to nodes on the ARB neighbor-joining tree. Signature sequences were detected using the ARB software package. A signature sequence is defined here as a short sequence that is present in a group of poribacterial sequences in a phylogenetic clade but is not found in any other clade in the poribacterial tree.Analysis of the 16S rRNA gene clone library sequences generated from sponge tissues revealed the presence of poribacteria in sponge individuals belonging to the orders Verongida, Astrophorida, Dictyoceratida, Haplosclerida, Lithistida, and Homosclerophorida, while poribacteria could not be detected in sponges belonging to the orders Hadromerida and Agelasida. In the order Halichondrida, poribacteria were detected in Xestospongia muta but not in Haliclona sp. Altogether, nine sponge species were added to the list of Poribacteria-containing sponges (Table (Table1).1). Three distinct clades were observed that were clearly supported by bootstrap values greater than 75 with every tree-building algorithm applied (Fig. (Fig.1),1), and one clade (clade I) was supported by bootstrap values of 64, 98, and 71 in distance, maximum parsimony, and maximum likelihood trees, respectively. Similarity calculations using approximately 740-bp amplified poribacterial 16S rRNA gene fragments and other poribacterial sequences from the NCBI database showed that the dissimilarity between clades was consistent with their separation in phylogenetic trees. For example, the levels of dissimilarity between members of clade I and clade II were 3 to 8%, while the levels of dissimilarity between members of clades I and III and between members of clades I and IV were 10 to 14% and 11 to 15%, respectively.Open in a separate windowFIG. 1.Neighbor-joining phylogenetic tree for poribacterial clones based on Poribacteria-specific PCR products (740 bp) of the 16S rRNA gene, showing relationships of poribacterial clones from different global regions. The poribacterial clones on the right are additional clones belonging to the same clades as strains in the tree at the same level. Bootstrap confidence values of >75% for distance, maximum parsimony, and maximum likelihood algorithm analyses are indicated by filled circles at nodes, and open circles indicate unsupported nodes. Prefixes for clones: A, Aplysina aerophoba; C, Aplysina cavernicola; F, Aplysina fistularis; L, Aplysina lacunosa; S, Ircinia sp.; P, Plakortis sp.; PC, Pseudoceratina clavata; RG, Rhabdastrella globostellata; T, Theonella swinhoei; X, Xestospongia muta. Scale bar = 0.1 nucleotide substitution per site.

TABLE 1.

Distribution of poribacteria in different demosponge orders
Sponge species or seawaterOrderGeographic locationaPresence of poribacteriabReference
Aplysina aerophobaVerongidaMED+2
Aplysina lacunosaVerongidaBAH+2
Aplysina fistularisVerongidaEPAC or BAH+2
Aplysina insularisVerongidaBAH+2
Verongula giganteaVerongidaBAH+2
Smenospongia aureaDictyoceratidaBAH+2
Aplysina cauliformisVerongidaBAH+This study
Aplysina archeriVerongidaBAH+This study
Aplysina cavernicolaVerongidaMED+This study
Pseudoceratina clavataVerongidaWPAC+This study
Rhabdastrella globostellataAstrophoridaWPAC+This study
Ircinia sp.DictyoceratidaBAH+This study
Xestospongia mutaHaploscleridaBAH+This study
Theonella swinhoeiLithistidaEPAC+This study
Plakortis sp.HomosclerophoridaBAH+This study
Chondrilla nuculaHadromeridaBAH2
Agelas wiedenmayeriAgelasidaBAH2
Agelas cerebrumAgelasidaBAHThis study
Axinella polypoidesHalichondridaMEDThis study
Ptilocaulis sp.HalichondridaBAH2
Dysidea avaraDictyoceratidaMEDThis study
Haliclona sp.HaploscleridaMEDThis study
Ectyoplasia feroxPoeciloscleridaBAH2
SeawaterNAcMEDThis study
Open in a separate windowaMED, Mediterranean Sea; BAH, Bahamas; WPAC, western Pacific Ocean; EPAC, eastern Pacific Ocean.bThe presence of poribacteria was evaluated by sequencing and phylogenetic analysis of amplified PCR products. +, present; −, absent.cNA, not applicable.Within each clade in the phylum Poribacteria, there were higher similarity values, including 94 to 100% among members of clade I, 94 to 99% among members of clade II, 96 to 99% among members of clade III, and 96 to 99% among members of clade IV. When members of the the phylum Poribacteria were compared to members of the Planctomycetes (Fig. (Fig.1),1), the 16S rRNA genes exhibited levels of sequence dissimilarity of up to 38%, consistent with the conclusion of Fieseler et al. concerning the separate phylum level status of poribacteria based on a similarity value of <75%. A phylogenetic correlation between sponge phylogeny and poribacterial phylogeny is not evident, since, for example, clones from A. fistularis and Aplysina aerophoba occurred in both clade I and clade II and one clone from A. aerophoba also occurred in clade III, while clones from P. clavata and R. globostellata occurred in clades I, II, and III but not in clade IV. Clades I and II included poribacterial clones derived from all sponge species occurring in all of the widely separated geographic regions examined in this study (Fig. (Fig.2).2). Clade III represented poribacterial clones derived from sponge species obtained in the eastern Pacific region, GBR, and the Bahamas but not in the Mediterranean region. The majority of poribacterial clones in clade IV were derived from sponge species obtained in the Bahamas, and one clade IV clone was obtained from a sponge species collected in the Mediterranean region.Open in a separate windowFIG. 2.Neighbor-joining phylogenetic tree for poribacterial clones based on Poribacteria-specific PCR products (740 bp) of the 16S rRNA gene, showing the internal relationships of and occurrence of clade I members in distinct sponge species representing cosmopolitan geographic regions. For an explanation of the colors, see Fig. Fig.1.1. Bootstrap confidence values of >75% for distance, maximum parsimony, and maximum likelihood algorithm analyses are indicated by filled circles at nodes, and open circles indicate unsupported nodes. Prefixes for clones: A, Aplysina aerophoba; C, Aplysina cavernicola; F, Aplysina fistularis; L, Aplysina lacunosa; S, Ircinia sp.; P, Plakortis sp.; PC, Pseudoceratina clavata; RG, Rhabdastrella globostellata; T, Theonella swinhoei; X, Xestospongia muta. Scale bar = 0.1 nucleotide substitution per site. Clones PC15, L8, T6, C2, P3, S2, and X1 were removed from this analysis to allow better branch resolution.Poribacterial clones from different sponges from widely separated marine habitats belonged to at least four major clades with similarities ranging from 94 and 96%. For clade III (Fig. (Fig.1),1), we detected a signature sequence characteristic of poribacterial clones from the GBR sponges R. globostellata and P. clavata. This signature sequence (CCA GTT AGC TTG ACG GTA) (Table (Table2)2) at E. coli positions 469 to 487 targeted 10 sequences, 5 of which were from GBR marine sponges generated in this study (clones RG68, RG112, RG105, PC96, and PC8). Another five poribacterial sequences were detected in an unpublished study investigating the microbial diversity in GBR sponges. This signature sequence indicates a specific geographic presence of poribacteria belonging to clade IV in the GBR region. In addition, a sequence (GAG TGT GAA ATG GCT TGG at E. coli positions 599 to 617) characteristic of clade IV was found in 11 sequences derived from sponges from the Bahamas and one sequence (A7) from a Mediterranean sponge.

TABLE 2.

Poribacterial signature sequences for clades III and IV, including a GBR-specific signature sequence (pori_SSIII) and a signature sequence specific to 11 of 12 sequences from the Bahamas (pori_SSIV)
Signature sequenceNameFull nameaE. coli positionSequenceb
pori_SSIIIPla101PPla101P*469GGUGAUAAG-==================-CCAUAGUA
Pla131PPla131P*469GGUGAUAAG-==================-CCAUAGUA
Pla134PPla134P*469GGUGAUAAG-==================-CCAUAGUA
Pla50PPla50P*469GGUGAUAAG-==================-CCAUAGUA
Pla82PPla82P*469GGUGAUAAG-==================-CCAUAGUA
PO68Pori clone RGPo68469GGUGAUAAG-==================-GAGAAAAG
PO112Pori clone RGPo112469GGUGAUAAU-==================-CCAUAGUA
PO105Pori clone RGPo105469GGUGAUAAG-==================-CCAUAGUA
PO96Uncultured Pori clone469GGUGAUAAG-==================-CCAUAGUA
PCPO8Pori clone PCPo8469GGUGAUAAG-==================-CCAUAGUA
pori_SSIVAY485286Uncultured Pori clone599ACAUUAGUC-==================-CUCAACCA
AY485285Uncultured Pori clone599ACAUNAGUC-==================-CUCAACNA
AY485284Uncultured Pori clone599ACAUUAGUC-==================-CUCAACCA
AY485281Uncultured Pori bacterium599ACAUUAGUC-==================-CUCAACCA
A7A7599AUAUUAGUC-==================-CUCAACCA
F2F2599ACAUAAGUC-==================-CUCAACCA
L16L16599ACAUUAGUC-==================-CUCAACCA
P20P20599ACAUAAGUC-==================-CUCAACCA
P38P38599ACAUUAGUC-==================-CUCAACCA
S6S6599AUAUUAGUC-==================-CUCAACCA
S10S10599AUAUUAGUC-==================-CUCAACCA
X18X18599ACAUUAGUC-==================-CUCAACCA
Open in a separate windowaAsterisks indicate poribacterial clones derived from the GBR sponge R. globostellata in a separate study.bThe internal sequence (indicated by equals signs) of each pori_SSIII clone is CCAGUUAGCUUGACGGUA, and that of each pori_SSIV clone is GAGUGUGAAAUGGCUUGG.Based on the data presented here, Poribacteria appears to be a bacterial phylum that is specifically found in several demosponge genera of the phylum Porifera (Table (Table1).1). To our knowledge, this is the only case of a bacterial phylum specifically associated with a marine invertebrate phylum. Certain phylum members appear to be widely distributed among sponges belonging to different species and in different geographic regions, forming sponge-specific lineages (3), but these are individual species level or at most genus level clades in a subdivision of a phylum rather than in a whole phylum.PCR analyses of seawater samples collected in this study (Table (Table1)1) and searches using nucleotide sequence databases of seawater metagenomes were negative for poribacteria. This is consistent with the concept that Poribacteria is a sponge-specific phylum. Within the sponges poribacteria are distributed among members of distinct demosponge orders that occur in various geographic locations, indicating that there is wide distribution of poribacteria among marine demosponges. Very similar 16S rRNA clone sequences that cluster in clade I were found in sponges from all geographic regions sampled in this study, including locations in the Northern and Southern hemispheres (Fig. (Fig.2).2). Similarly, clade II contains poribacterial clones from the Mediterranean Aplysina species and from GBR Pseudoceratina and Rhabdastrella species. This appears to contrast, albeit at a lower level of resolution, with results suggesting that bacterial populations are endemic in different geographic regions, e.g., with the findings that marine bacterioplankton communities include few cosmopolitan operational taxonomic units (8), that fluorescent Pseudomonas genotypes from soil are endemic at different geographic sites (1), and that hyperthermophilic Sulfolobus archaea from different geothermal areas are genetically divergent (12). Judging the endemicity of populations in different geographic regions may depend on the taxonomic scale used to distinguish populations (1). In this study we provide evidence that at least some clades may be relatively characteristic of particular regions, e.g., GBR clade III (Table (Table2).2). It is remarkable that in the case of the sponge species R. globostellata and P. clavata from a single geographic region (GBR), the microbial communities include representatives of distantly related poribacterial clades II and III, whose sequences exhibit levels of dissimilarity ranging from 10 to 13%. In another case poribacteria belonging to clades I, II, and IV were found in a single host, A. aerophoba, from the Mediterranean. Thus, members of widely divergent poribacterial clades occur in the same specimen in sponges in widely separated geographic regions in the world''s oceans. Three different sponge species belonging to three different genera exhibit this phenomenon.The morphology and life strategy of sponges have remained unchanged for the past 580 million years, as judged by the dramatic similarity of the morphologies of Precambrian fossils to the morphologies of recent sponges (6). Adaptation of the poribacteria to this niche might have taken place early in evolution before the various sponge orders separated from each other. It seems likely that poribacteria diverged from other bacterial phyla long before evolution of the metazoans as part of the fan-like radiation by which all bacterial phyla appear to have arisen (4). This bacterial radiation may have resulted in the divergence of the clades that we have observed for the poribacteria, but there is no indication of cospeciation between host sponges and the poribacteria.In summary, poribacteria exhibit considerable diversity and are classified into four phylogenetic clades. Poribacteria seem to be widely distributed among many different marine demosponge genera, and further studies are needed to explain the nature of the poribacterium-sponge interaction.  相似文献   
998.
999.

Background

The methods used for sample selection and processing can have a strong influence on the expression values obtained through microarray profiling. Laser capture microdissection (LCM) provides higher specificity in the selection of target cells compared to traditional bulk tissue selection methods, but at an increased processing cost. The benefit gained from the higher tissue specificity realized through LCM sampling is evaluated in this study through a comparison of microarray expression profiles obtained from same-samples using bulk and LCM processing.

Methods

Expression data from ten lung adenocarcinoma samples and six adjacent normal samples were acquired using LCM and bulk sampling methods. Expression values were evaluated for correlation between sample processing methods, as well as for bias introduced by the additional linear amplification required for LCM sample profiling.

Results

The direct comparison of expression values obtained from the bulk and LCM sampled datasets reveals a large number of probesets with significantly varied expression. Many of these variations were shown to be related to bias arising from the process of linear amplification, which is required for LCM sample preparation. A comparison of differentially expressed genes (cancer vs. normal) selected in the bulk and LCM datasets also showed substantial differences. There were more than twice as many down-regulated probesets identified in the LCM data than identified in the bulk data. Controlling for the previously identified amplification bias did not have a substantial impact on the differences identified in the differentially expressed probesets found in the bulk and LCM samples.

Conclusion

LCM-coupled microarray expression profiling was shown to uniquely identify a large number of differentially expressed probesets not otherwise found using bulk tissue sampling. The information gain realized from the LCM sampling was limited to differential analysis, as the absolute expression values obtained for some probesets using this study's protocol were biased during the second round of amplification. Consequently, LCM may enable investigators to obtain additional information in microarray studies not easily found using bulk tissue samples, but it is of critical importance that potential amplification biases are controlled for.  相似文献   
1000.
The prevalence of obesity in children and adults in the United States has increased dramatically over the past decade. Besides environmental factors, genetic factors are known to play an important role in the pathogenesis of obesity. A number of genetic determinants of adult BMI have already been established through genome‐wide association (GWA) studies. In this study, we examined 25 single‐nucleotide polymorphisms (SNPs) corresponding to 13 previously reported genomic loci in 6,078 children with measures of BMI. Fifteen of these SNPs yielded at least nominally significant association to BMI, representing nine different loci including INSIG2, FTO, MC4R, TMEM18, GNPDA2, NEGR1, BDNF, KCTD15, and 1q25. Other loci revealed no evidence for association, namely at MTCH2, SH2B1, 12q13, and 3q27. For the 15 associated variants, the genotype score explained 1.12% of the total variation for BMI z‐score. We conclude that among 13 loci that have been reported to associate with adult BMI, at least nine also contribute to the determination of BMI in childhood as demonstrated by their associations in our pediatric cohort.  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