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 62 毫秒
1.
GB virus C/hepatitis G (GBV-C) is an RNA virus of the family Flaviviridae. Despite replicating with an RNA-dependent RNA polymerase, some previous estimates of rates of evolutionary change in GBV-C suggest that it fixes mutations at the anomalously low rate of ∼10−7 nucleotide substitution per site, per year. However, these estimates were largely based on the assumption that GBV-C and its close relative GBV-A (New World monkey GB viruses) codiverged with their primate hosts over millions of years. Herein, we estimated the substitution rate of GBV-C using the largest set of dated GBV-C isolates compiled to date and a Bayesian coalescent approach that utilizes the year of sampling and so is independent of the assumption of codivergence. This revealed a rate of evolutionary change approximately four orders of magnitude higher than that estimated previously, in the range of 10−2 to 10−3 sub/site/year, and hence in line with those previously determined for RNA viruses in general and the Flaviviridae in particular. In addition, we tested the assumption of host-virus codivergence in GBV-A by performing a reconciliation analysis of host and virus phylogenies. Strikingly, we found no statistical evidence for host-virus codivergence in GBV-A, indicating that substitution rates in the GB viruses should not be estimated from host divergence times.  相似文献   

2.
Bluetongue virus (BTV) is a midge-borne member of the genus Orbivirus that causes an eponymous debilitating livestock disease of great agricultural impact and which has expanded into Europe in recent decades. Reassortment among the ten segments comprising the double-stranded (ds) RNA genome of BTV has played an important role in generating the epidemic strains of this virus in Europe. In this study, we investigated the dynamics of BTV genome segment evolution utilizing time-structured data sets of complete sequences from four segments, totalling 290 sequences largely sampled from ruminant hosts. Our analysis revealed that BTV genome segments generally evolve under strong purifying selection and at substitution rates that are generally lower (mean rates of ~0.5–7 × 10−4 nucleotide substitutions per site, per year) than vector-borne positive-sense viruses with single-strand (ss) RNA genomes. These also represent the most robust estimates of the nucleotide substitution rate in a dsRNA virus generated to date. Additionally, we determined that patterns of geographic structure and times to most recent common ancestor differ substantially between each segment, including a relatively recent origin for the diversity of segment 10 within the past millennium. Together, these findings demonstrate the effect of reassortment to decouple the evolutionary dynamics of BTV genome segments.  相似文献   

3.
Hantaviruses are rodent-borne Bunyaviruses that infect the Arvicolinae, Murinae, and Sigmodontinae subfamilies of Muridae. The rate of molecular evolution in the hantaviruses has been previously estimated at approximately 10(-7) nucleotide substitutions per site, per year (substitutions/site/year), based on the assumption of codivergence and hence shared divergence times with their rodent hosts. If substantiated, this would make the hantaviruses among the slowest evolving of all RNA viruses. However, as hantaviruses replicate with an RNA-dependent RNA polymerase, with error rates in the region of one mutation per genome replication, this low rate of nucleotide substitution is anomalous. Here, we use a Bayesian coalescent approach to estimate the rate of nucleotide substitution from serially sampled gene sequence data for hantaviruses known to infect each of the 3 rodent subfamilies: Araraquara virus (Sigmodontinae), Dobrava virus (Murinae), Puumala virus (Arvicolinae), and Tula virus (Arvicolinae). Our results reveal that hantaviruses exhibit short-term substitution rates of 10(-2) to 10(-4) substitutions/site/year and so are within the range exhibited by other RNA viruses. The disparity between this substitution rate and that estimated assuming rodent-hantavirus codivergence suggests that the codivergence hypothesis may need to be reevaluated.  相似文献   

4.
Viral abundance, burst sizes, lytic production and temperate phage were investigated in land-fast ice at two sites in Prydz Bay Antarctica (68°S, 77°E) between April and November 2008. Both ice cores and brine were collected. There was no seasonal pattern in viral or bacterial numbers. Across the two sites virus abundances ranged between 0.5 × 105 and 5.1 × 105 viruses ml−1 in melted ice cores and 0.6 × 105–3.5 × 105 viruses ml−1 in brine, and bacterial abundances between 2.7 × 104 and 17.3 × 104 cells ml−1 in melted ice cores and 3.9 × 104–32.5 × 104 cells ml−1 in brine. Virus to bacterium ratios (VBR) showed a clear seasonal pattern in ice cores with lowest values in winter (range 1.2–20.8), while VBRs in brine were lower (0.2–4.9). Lytic viral production range from undetectable to 2.0 × 104 viruses ml−1 h−1 in ice cores with maximum rates in September and November. In brine maximum, lytic viral production occurred in November (1.18 × 104 viruses ml−1 h−1). Low burst sizes were typical (3.94–4.03 viruses per bacterium in ice cores and 3.16–4.0 viruses per bacterium in brine) with unusually high levels of visibly infected cells—range 40–50%. This long-term investigation revealed that viral activity was apparent within the sea ice throughout its annual cycle. The findings are discussed within the context of limited data available on viruses in sea ice.  相似文献   

5.
Molecular evolution of nitrate reductase genes   总被引:9,自引:0,他引:9  
To understand the evolutionary mechanisms and relationships of nitrate reductases (NRs), the nucleotide sequences encoding 19 nitrate reductase (NR) genes from 16 species of fungi, algae, and higher plants were analyzed. The NR genes examined show substantial sequence similarity, particularly within functional domains, and large variations in GC content at the third codon position and intron number. The intron positions were different between the fungi and plants, but conserved within these groups. The overall and nonsynonymous substitution rates among fungi, algae, and higher plants were estimated to be 4.33 × 10−10 and 3.29 × 10−10 substitutions per site per year. The three functional domains of NR genes evolved at about one-third of the rate of the N-terminal and the two hinge regions connecting the functional domains. Relative rate tests suggested that the nonsynonymous substitution rates were constant among different lineages, while the overall nucleotide substitution rates varied between some lineages. The phylogenetic trees based on NR genes correspond well with the phylogeny of the organisms determined from systematics and other molecular studies. Based on the nonsynonymous substitution rate, the divergence time of monocots and dicots was estimated to be about 340 Myr when the fungi–plant or algae–higher plant divergence times were used as reference points and 191 Myr when the rice–barley divergence time was used as a reference point. These two estimates are consistent with other estimates of divergence times based on these reference points. The lack of consistency between these two values appears to be due to the uncertainty of the reference times. Received: 10 April 1995 / Accepted: 10 September 1995  相似文献   

6.
Paramyxoviruses are responsible for considerable disease burden in human and wildlife populations: measles and mumps continue to affect the health of children worldwide, while canine distemper virus causes serious morbidity and mortality in a wide range of mammalian species. Although these viruses have been studied extensively at both the epidemiological and the phylogenetic scales, little has been done to integrate these two types of data. Using a Bayesian coalescent approach, we infer the evolutionary and epidemiological dynamics of measles, mumps and canine distemper viruses. Our analysis yielded data on viral substitution rates, the time to common ancestry, and elements of their demographic history. Estimates of rates of evolutionary change were similar to those observed in other RNA viruses, ranging from 6.585 to 11.350 × 10−4 nucleotide substitutions per site, per year. Strikingly, the mean Time to the Most Recent Common Ancestor (TMRCA) was both similar and very recent among the viruses studied, ranging from only 58 to 91 years (1908 to 1943). Worldwide, the paramyxoviruses studied here have maintained a relatively constant level of genetic diversity. However, detailed heterchronous samples illustrate more complex dynamics in some epidemic populations, and the relatively low levels of genetic diversity (population size) in all three viruses is likely to reflect the population bottlenecks that follow recurrent outbreaks.  相似文献   

7.
Databases on effects of chronic low-LET radiation exposure were analyzed by non-parametric statistical methods, to estimate the threshold dose rates above which radiation effects can be expected in vertebrate organisms. Data were grouped under three umbrella endpoints: effects on morbidity, reproduction, and life shortening. The data sets were compiled on a simple ‘yes’ or ‘no’ basis. Each data set included dose rates at which effects were reported without further details about the size or peculiarity of the effects. In total, the data sets include 84 values for endpoint “morbidity”, 77 values for reproduction, and 41 values for life shortening. The dose rates in each set were ranked from low to higher values. The threshold TDR5 for radiation effects of a given umbrella type was estimated as a dose rate below which only a small percentage (5%) of data reported statistically significant radiation effects. The statistical treatment of the data sets was performed using non-parametric order statistics, and the bootstrap method. The resulting thresholds estimated by the order statistics are for morbidity effects 8.1 × 10−4 Gy day−1 (2.0 × 10−4–1.0 × 10−3), reproduction effects 6.0 × 10−4 Gy day−1 (4.0 × 10−4–1.5 × 10−3), and life shortening 3.0 × 10−3 Gy day−1 (1.0 × 10−3–6.0 × 10−3), respectively. The bootstrap method gave slightly lower values: 2.1 × 10−4 Gy day−1 (1.4 × 10−4–3.2 × 10−4) (morbidity), 4.1 × 10−4 Gy day−1 (3.0 × 10−4–5.7 × 10−4) (reproduction), and 1.1 × 10−3 Gy day−1 (7.9 × 10−4–1.3 × 10−3) (life shortening), respectively. The generic threshold dose rate (based on all umbrella types of effects) was estimated at 1.0 × 10−3 Gy day−1.  相似文献   

8.
A field study was conducted in a nutrient-impacted marsh in Water Conservation Area 2A (WCA-2A) of the Everglades in southern Florida, USA, to evaluate early stages of plant litter (detritus) decomposition along a well-documented trophic gradient, and to determine the relative importance of environmental factors and substrate composition in governing decomposition rate. Vertically stratified decomposition chambers containing native plant litter (cattail and sawgrass leaves) were placed in the soil and water column along a 10-km transect coinciding with a gradient of soil phosphorus (P) enrichment. Decomposition rate varied significantly along the vertical water–soil profile, with rates typically higher in the water column and litter layer than below the soil surface, presumably in response to vertical gradients of such environmental factors as O2 and nutrient availability. An overall decrease in decomposition rate occurred along the soil P gradient (from high- to low-impact). First-order rate constant (k) values for decomposition ranged from 1.0 to 9.2 × 10−3 day−1 (mean = 2.8 ×10−3 day−1) for cattails, and from 6.7 × 10−4 to 3.0 ×  10−3 day−1 (mean = 1.7 ×  10−3 day−1) for sawgrass. Substantial N and P immobilization occurred within the litter layer, being most pronounced at nutrient-impacted sites. Nutrient content of the decomposing plant tissue was more strongly correlated to decomposition rate than was the nutrient content of the surrounding soil and water. Our experimental results suggest that, although decomposition rate was significantly affected by initial substrate composition, the external supply or availability of nutrients probably played a greater role in controlling decomposition rate. It was also evident that nutrient availability for litter decomposition was not accurately reflected by ambient nutrient concentration, e.g., water and soil porewater nutrient concentration.  相似文献   

9.
Seven new Italian and two new British HTLV-II isolates were obtained from injecting drug users and the entire long terminal repeat (LTR) region was sequenced. Restriction analysis showed that all the Italian isolates are of the IIb subtype, whereas the British isolates are of the IIa subtype. To understand whether the further differentiation of each two principal HTLV-II subtypes in several subgroups could be statistically supported by phylogenetic analysis, the neighbor-joining, parsimony, and maximum likelihood methods were used. The separation between IIa and IIb is very well supported by all three methods. At least two phylogenetic subgroups exist within the HTLV-IIa and at least three within the HTLV-IIb subtype. In the present analysis, no statistical support was obtained for additional phylogroups. Two particular subgroups seem interesting because they include all European and North American injecting drug user strains within the IIa and IIb subtypes, respectively. These data confirm that European HTLV-II infection among drug users is probably derived from North America. They also suggest that though a certain differentiation by restriction analysis in different subgroups is possible, carefully interpreted phylogenetic analyses remain necessary. Using the likelihood ratio test, a molecular clock for the drug user strains was calibrated. A fixation rate between 1.08 × 10−4 and 2.7 × 10−5 nucleotide substitutions per site per year was calculated for the IIa and IIb injecting drug user strains. This is the lowest fixation rate so far reported for RNA viruses, including for HIV, which typically range between 10−2 and 10−4.  相似文献   

10.
Molecular evolutionary analyses were carried out to elucidate the phylogenetic relationships, the evolutionary rate, and the divergence times of hepatitis C viruses. Using the nucleotide sequences of the viruses isolated from various locations in the world, we constructed phylogenetic trees. The trees showed that strains isolated from a single location were not necessarily clustered as a group. This suggests that the viruses may be transferred with blood on a worldwide scale. We estimated the evolutionary rates at synonymous and nonsynonymous sites for all genes in the viral genome. We then found that the rate (1.35 × 10–3 per site per year) at synonymous sites for the C gene was much smaller than those for the other genes (e.g., 6.29 × 10–3 per site per year for the E gene). This indicates that a special type of functional constraint on synonymous substitutions may exist in the C gene. Because we found an open reading frame (ORF) with the C gene region, the possibility exists that synonymous substitutions for the C gene are constrained by the overlapping ORF whose reading frame is different from that of the C gene. Applying the evolutionary rates to the trees, we also suggest that major groups of hepatitis C viruses diverged from their common ancestor several hundred years ago. Correspondence to: T. Gojobori  相似文献   

11.
Issues associated with newly emerging viruses, their genetic diversity, and viral evolution in modern environments are currently attracting growing attention. In this study, a phylogenetic analysis was performed and the evolution rate was evaluated for such pathogenic flaviviruses endemic to Russia as tick-borne encephalitis virus (TBEV) and Powassan virus (PV). The analysis involved 47 nucleotide sequences of the TBEV genome region encoding protein E and 17 sequences of the PV NS5-encoding region. The nucleotide substitution rate was estimated as 1.4 × 10−4 and 5.4 × 10−5 substitutions per site per year for the E protein-encoding region of the TBEV genome and for the NS5 genome region of PV, respectively. The ratio of non-synonymous to synonymous nucleotide substitutions (dN/dS) in viral sequences was calculated as 0.049 for TBEV and 0.098 for PV. The highest dN/dS values of 0.201–0.220 were found in the subcluster of Russian and Canadian PV strains, and the lowest value of 0.024 was observed in the cluster of Russian and Chinese strains of the Far Eastern TBEV genotype. Evaluation of time intervals between the events of viral evolution showed that the European subtype of TBEV diverged from the common TBEV ancestor approximately 2750 years ago, while the Siberian and Far Eastern subtypes emerged approximately 2250 years ago. The PV was introduced into its natural foci of the Russian Primorskii krai only approximately 70 years ago; these strains were very close to Canadian PV strains. The pattern of PV evolution in North America was similar to the evolution of the Siberian and Far Eastern TBEV subtypes in Asia. The moments of divergence between major genetic groups of TBEV and PV coincide with historical periods of climate warming and cooling, suggesting that climate change was a key factor in the evolution of flaviviruses in past millennia.  相似文献   

12.
The effects of single pulses and multiple pulses of 7 MV electrons on micronuclei (MN) induction in cytokinesis-blocked human peripheral blood lymphocytes (PBLs) were investigated over a wide range of dose rates per pulse (instantaneous dose rate). PBLs were exposed to graded doses of 2, 3, 4, 6, and 8 Gy of single electron pulses of varying pulse widths at different dose rates per pulse, ranging from 1 × 106 Gy s−1 to 3.2 × 108 Gy s−1. Different dose rates per pulse were achieved by changing the dose per electron pulse by adjusting the beam current and pulse width. MN yields per unit absorbed dose after irradiation with single electron pulses were compared with those of multiple pulses of electrons. A significant decrease in the MN yield with increasing dose rates per pulse was observed, when dose was delivered by a single electron pulse. However, no reduction in the MN yield was observed when dose was delivered by multiple pulses of electrons. The decrease in the yield at high dose rates per pulse suggests possible radical recombination, which leads to decreased biological damage. Cellular response to the presence of very large numbers of chromosomal breaks may also alter the damage.  相似文献   

13.
Tobamoviruses, mostly isolated from solanaceous plants, may represent ancient virus lineages that have codiverged with their hosts. Recently completed nucleotide sequences of six nonsolanaceous tobamoviruses allowed assessment of the codivergence hypothesis and support a third subgroup within tobamoviruses. The genomic sequences of 12 tobamoviruses and the partial sequences of 11 others have been analyzed. Comparisons of the predicted protein sequences revealed three clusters of tobamoviruses, corresponding to those infecting solanaceous species (subgroup 1), those infecting cucurbits and legumes (subgroup 2), and those infecting crucifers. The orchid-infecting odontoglossum ringspot tobamovirus was associated with subgroup 1 genomes by its coat and movement protein sequences, but with the crucifer-pathogenic tobamoviruses by the remainder of its genome, suggesting that it is the progeny of a recombinant. For four of five genomic regions, subgroup 1 and 3 genomes were equidistant from a subgroup 2 genome chosen for comparison, suggesting uniform rates of evolution. A phylogenetic tree of plant families based on the tobamoviruses they harbor was congruent with that based on rubisco sequences but had a different root, suggesting that codivergence was tempered by rare events of viruses of one family colonizing another family. The proposed subgroup 3 viruses probably have an origin of virion assembly in the movement protein gene, a large (25-codon) overlap of movement and coat protein open reading frames, and a comparably shorter genome. Codon-position- dependent base compositions and codon prevalences suggested that the coat protein frame of the overlap region was ancestral. Bootstrapped parsimony analysis of the nucleotides in the overlap region and of the sequences translated from the -1 frame (the subgroup 3 movement protein frame) of this region produced trees inconsistent with those deduced from other regions. The results are consistent with a model in which a no or short overlap organization was ancestral. Despite encoding of subgroup 2 and 3 movement protein C-termini by nonhomologous nucleotides, weak similarities between their amino acid sequences suggested convergent sequence evolution.   相似文献   

14.
With the aim of elucidating evolutionary features of GB virus C/hepatitis G virus (GBV-C/HGV), molecular evolutionary analyses were conducted using the entire coding region of this virus. In particular, the rate of nucleotide substitution for this virus was estimated to be less than 9.0 × 10−6 per site per year, which was much slower than those for other RNA viruses. The phylogenetic tree reconstructed for GBV-C/HGV, by using GB virus A (GBV-A) as outgroup, indicated that there were three major clusters (the HG, GB, and Asian types) in GBV-C/HGV, and the divergence between the ancestor of GB- and Asian-type strains and that of HG-type strains first took place more than 7000–10,000 years ago. The slow evolutionary rate for GBV-C/HGV suggested that this virus cannot escape from the immune response of the host by means of producing escape mutants, implying that it may have evolved other systems for persistent infection. Received: 2 June 1998 / Accepted: 8 August 1998  相似文献   

15.
Viruses play a significant role in nutrient cycling within the world’s oceans and are important agents of horizontal gene transfer, but little is know about their entrainment into sea ice or their temporal dynamics once entrained. Nilas, grease ice, pancake ice, first-year sea ice floes up to 78 cm in thickness, and under-ice seawater were sampled widely across Amundsen Gulf (ca. 71° N, 125° W71^\circ \hbox{N}, 125^\circ \hbox{W}) for concentrations of viruses and bacteria. Here, we report exceptionally high virus-to-bacteria ratios in seawater (45–340) and sea ice (93–2,820) during the autumn freeze-up. Virus concentrations ranged from 4.8 to 27 × 106  ml−1 in seawater and, scaled to brine volume, 5.5 to 170 × 107 ml−1 in sea ice. Large enrichment indices indicated processes of active entrainment from source seawater, or viral production within the ice, which was observed in 2 of 3 bottle incubations of sea ice brine at a temperature (-7°C-7^\circ\hbox{C}) and salinity ( 110 \permille110 \permille) approximating that in situ. Median predicted virus-to-bacteria contact rates (relative to underlying seawater) were greatest in the top of thick sea ice (66–78 cm: 130×) and lowest in the bottom of medium-thickness ice (33–37 cm: 23×). The great abundance of viruses and more frequent interactions between bacteria and viruses predicted in sea ice relative to underlying seawater suggest that sea ice may be a hot spot for virally mediated horizontal gene transfer in the polar marine environment.  相似文献   

16.
Kim JJ  Kim HH  Park JH  Ryu HJ  Kim J  Moon S  Gu H  Kim HT  Lee JY  Han BG  Park C  Kimm K  Park CS  Lee JK  Oh B 《Immunogenetics》2005,57(9):636-643
Asthma is a chronic inflammatory disorder of the airways, and a number of genetic loci are associated with the disease. Candidate gene association studies have been regarded as effective tools to study complex traits. Knowledge of the sequence variation and structure of the candidate genes is required for association studies. Thus, we investigated the genetic variants of 32 asthma candidate genes selected by colocalization of positional and functional candidate genes. We screened all exons and promoter regions of those genes using 12 healthy individuals and 12 asthma patients and identified a total of 418 single nucleotide polymorphisms (SNPs), including 270 known SNPs and 148 novel SNPs. Levels of nucleotide diversity varied from gene to gene (0.72×10−4–14.53×10−4), but the average nucleotide diversity between coding SNPs (cSNPs) and noncoding SNPs was roughly equivalent (4.63×10−4 vs 4.69×10−4). However, nucleotide diversity of cSNPs was strongly correlated to codon degeneracy. Nucleotide diversity was much higher at fourfold degenerate sites than at nondegenerate sites (9.42×10−4 vs 3.14×10−4). Gene-based haplotype analysis of asthma-associated genes in this study revealed that common haplotypes (frequency >5%) represented 90.5% of chromosomes, and they could be uniquely identified with five or fewer haplotype-tagging SNPs per gene. Therefore, our results may have important implications for the selection of asthma candidate genes and SNP markers for comprehensive association studies using large sample populations.  相似文献   

17.
Woo J  Kim Y  Lee C 《Molecular biology reports》2012,39(3):2259-2265
We examined genetic associations of previously identified sequence variants with bone mineral density and their heterogeneity by gender. Large-scale cohort data were used including a total of 8,419 subjects (4,034 males and 4,385 females) from the Korean Association REsource (KARE) cohort. Bone speed of sound (SOS) values were measured at distal radius or mid-shaft tibia by quantitative ultrasound. Genotypic associations of SOS were tested with each of nucleotide sequence variants identified by previous studies. The genetic association analysis revealed that 2 out of 11 nucleotide sequence variants were associated with SOS (rs1721400, rs7776725, P < 7.58 × 10−4). Further analysis with partitioning data by gender showed that the mid-shaft tibia phenotypes were associated with the rs1721400 and rs7776725 in females (P < 3.79 × 10−4), but not in males (P > 3.79 × 10−4). The current study suggested female-specific associations of rs1721400 and rs7776725 with bone mineral density and heterogeneity of genetic association by skeletal site measured for bone mineral density.  相似文献   

18.
Seashore paspalum (Paspalum vaginatum Swartz) is a salt tolerant, fine textured turfgrass used on golf courses in coastal, tropical, and subtropical regions. A callus induction and plant regeneration protocol for this commercially important turfgrass species has been developed. Induction of highly regenerable callus with approximately 400 shoots per cultured immature inflorescence (1 cm in length) was achieved by culturing 0.2 cm segments on media with 3 mg l−1 3,6-dichloro-2-methoxybenzoic acid (dicamba) and 0.1 or 1.0 mg l−1 benzylaminopurine (BA). A multifactorial experiment demonstrated the combination of 3 mg l−1 dicamba and 1.0 mg l−1 BA for induction of callus resulted in 12 times higher plant regeneration frequency compared to 3 mg l−1 2,4-dichlorophenoxyacetic acid (2,4-D) alone or ten times higher plant regeneration frequency than the combination of 3 mg l−1 2,4-D and 1.0 mg l−1 BA. These results are expected to support the development of a genetic transformation protocol for seashore paspalum.  相似文献   

19.
We investigated the effects of limiting (1.96 × 10−9 mol l−1 total Cu, corresponding to pCu 14.8; where pCu = −log [Cu2+]) and toxic Cu concentrations up to 8.0 × 10−5 mol l−1 total Cu (equivalent to pCu 9.5) on growth rates and photosynthetic activity of exponentially grown Phaeocystis cordata, using batch and semi-continuous cultures. With pulse amplitude modulated (PAM) fluorometry, we determined the photochemical response of P. cordata to the various Cu levels, and showed contrasting results for the batch and semi-continuous cultures. Although maximum photosystem II (PSII) quantum yield (ΦM) was optimal and constant in the semi-continuous P. cordata, the batch cultures showed a significant decrease in ΦM with culture age (0–72 h). The EC50 for the batch cultures was higher (2.0 × 10−10 mol l−1, pCu9.7), than that for the semi-continuous cultures (6.3 × 10−11 mol l−1, pCu10.2). The semi-continuous cultures exhibited a systematic and linear decrease in ΦM as Cu levels increased (for [Cu2+] < 1.0 × 10−12 mol l−1, pCu12.0), however, no effect of high Cu was observed on their operational PSII quantum yield (Φ′M). Similarly, semi-continuous cultures exhibited a significant decrease in ΦM, but not in Φ′M, because of low-Cu levels. Thus, Cu toxicity and Cu limitation damage the PSII reaction centers, but not the processes downstream of PSII. Quenching mechanisms (NPQ and Q n) were lower under high Cu relative to the controls, suggesting that toxic Cu impairs photo-protective mechanisms. PAM fluorometry is a sensitive tool for detecting minor physiological variations. However, culturing techniques (batch vs. semi-continuous) and sampling time might account for literature discrepancies on the effects of Cu on PSII. Semi-continuous culturing might be the most adequate technique to investigate Cu effects on PSII photochemistry.  相似文献   

20.
In a panel of seven genotypes, 437 expressed sequence tag (EST)-derived DNA fragments were sequenced. Single nucleotide polymorphisms (SNPs) that were polymorphic between the parents of three mapping populations were mapped by heteroduplex analysis and a genome-wide consensus map comprising 216 EST-derived SNPs and 4 InDel (insertion/deletion) markers was constructed. The average frequency of SNPs amounted to 1/130 bp and 1/107.8 bp for a set of randomly selected and a set of mapped ESTs, respectively. The calculated nucleotide diversities (π) ranged from 0 to 40.0 × 10−3 (average 3.1 × 10−3) and 0.52 × 10−3 to 39.51 × 10–3 (average 4.37 × 10−3) for random and mapped ESTs, respectively. The polymorphism information content value for mapped SNPs ranged from 0.24 to 0.50 with an average of 0.34. As expected, combination of SNPs present in an amplicon (haplotype) exhibited a higher information content ranging from 0.24 to 0.85 with an average of 0.50. Cleaved amplified polymorphic sequence assays (including InDels) were designed for a total of 87 (39.5%) SNP markers. The high abundance of SNPs in the barley genome provides avenues for the systematic development of saturated genetic maps and their integration with physical maps. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users. Both R. Kota and R.K. Varshney contributed equally to this work.  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