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 15 毫秒
1.
In tropical evergreen forest in the Kolli Hills of the Indian Eastern Ghats, four 2 ha (100 m × 200 m) replicate plots (two plots each in undisturbed and human-impacted sites), were inventoried for species diversity of lianas 5 cm girth at breast height (g.b.h.) and their relationships with 30 cm g.b.h. host trees. Liana diversity included 26 species from 18 families and 24 genera. The population density and basal area of lianas in the study plots were 48 individuals ha–1 and 0.23 m2 ha–1, respectively, while those of the trees were 478 stems ha–1 and 43.6 m2 ha–1, respectively. As the lianas and their hosts had often been cut in the disturbed sites, their diversity was less there than in the undisturbed sites. Five (19%) liana species were common to all four sites. Three lianas, Hiptage benghalensis (Malpighiaceae), Elaeagnus indica (Elaeagnaceae) and Gnetum ula (Gnetaceae) were dominant. The twining mechanism (54% of liana species and 71% of individuals) and zoochorous diaspores (73% of species and 77% of individuals) predominated. A total of 336 trees from 39 species, 34 genera and 22 families hosted 345 lianas. The ratio of liana : host for species was 1 : 1.5 and for individuals was 1 : 1. Liana preferences for certain host trees, host girth classes and trellis heights were evident.  相似文献   

2.
The sterically hindered tetrakis-(3-(p-tolylpyrazolyl)borate [pz0Tpp-Tol] has been prepared and its reaction with CuX2.nH2O (X = Cl or acetate (OAc), M(NO3)2.6H2O (M = Ni, or Co) and MCl2 (M = Zn or Cd) has been investigated. [M(pz0Tpp-Tol)X(Hpzp-Tol)] (M = Cu, X = Cl or OAc; M = Ni or Co, X = NO3) and [M(pz0Tpp-Tol)Cl(Hpzp-Tol)2-n(H2O)n] have been synthesised and their spectroscopic properties described, the five-coordinated Cu species being also structurally characterized. The methyl groups in the para-tolyl fragments of the ligand strongly influences the stoichiometry and structure of the metal complexes.  相似文献   

3.
Sphagna are vulnerable to enhanced nitrogen (N) deposition. This article reports how the green (shade, under Calluna) and red (open grown) Sphagnum capillifolium respond to ammonium and nitrate additions of 56 kg N ha−1 y−1 over the background of 8-10 kg N ha−1 y−1 on an ombrotrophic bog in the Scottish Borders after seven years. Samples and measurements were made during a range of hydrated and desiccated conditions in the summer of 2009. Both ammonium and nitrate increased moss N concentration, but while ammonium decreased cross-sectional area of leaf hyaline cells and the leaf hyaline/chlorophyllose cell area ratio, nitrate increased both of them and capitulum pH. The changes in leaf morphology have not previously been reported to our knowledge. Especially the red S. capillifolium was affected by ammonium with significant changes in shoot N concentration (+71%) and the cross-sectional area of leaf chlorophyllose cells (+67%), and reductions in shoot dry weight (−30%) and fresh weight (−42%), the cross-sectional area of leaf hyaline cells (−24%), the leaf hyaline/chlorophyllose cell area ratio (−54%), as well as in chlorophyll fluorescence (measured as Fv/Fm) of desiccated capitulum (−65%) (all p < 0.05). These observations show that N deposition may affect moss physiology also through changes in leaf anatomy and morphology. The results also highlight potential sampling issues and causes of variability in N responses when collecting variably pigmented Sphagna.  相似文献   

4.
Mammalian metallothioneins (MTs) are a family of small cysteine rich proteins believed to have a number of physiological functions, including both metal ion homeostasis and toxic metal detoxification. Mammalian MTs bind 7 Zn2+ or Cd2+ ions into two distinct domains: an N-terminal β-domain that binds 3 Zn2+ or Cd2+, and a C-terminal α-domain that binds 4 Zn2+ or Cd2+. Although stepwise metalation to the saturated M7-MT (where M = Zn2+ or Cd2+) species would be expected to take place via a noncooperative mechanism involving the 20 cysteine thiolate ligands, literature reports suggest a cooperative mechanism involving cluster formation prior to saturation of the protein. Electrospray ionization mass spectrometry (ESI-MS) provides this sensitivity through delineation of all species (Mn-MT, n = 0-7) coexisting at each step in the metalation process. We report modeled ESI-mass spectral data for the stepwise metalation of human recombinant MT 1a (rhMT) and its two isolated fractions for three mechanistic conditions: cooperative (where the binding affinities are: K1 < K2 < K3 < ··· < K7), weakly cooperative (where K1 = K2 = K3 = ··· = K7), and noncooperative, (where K1 > K2 > K3 > ··· > K7). Detailed ESI-MS metalation data of human recombinant MT 1a by Zn2+ and Cd2+ are also reported. Comparison of the experimental data with the predicted mass spectral data provides conclusive evidence that metalation occurs in a noncooperative fashion for Zn2+ and Cd2+ binding to rhMT 1a.  相似文献   

5.
The PPARγ agonist Rosiglitazone exerts anti-hyperglycaemic effects by regulating the long-term expression of genes involved in metabolism, differentiation and inflammation. In the present study, Rosiglitazone treatment rapidly inhibited (5-30 min) the ER Ca2+ ATPase SERCA2b in monocytic cells (IC50 = 1.88 μM; p < 0.05), thereby disrupting short-term Ca2+ homeostasis (resting [Ca2+]cyto = 121.2 ± 2.9% basal within 1 h; p < 0.05). However, extended Rosiglitazone treatment (72 h) induced dose-dependent SERCA2b up-regulation, and restored calcium homeostasis, in monocytic cells (SERCA2b mRNA: 138.7 ± 5.7% basal (1 μM)/215.0 ± 30.9% basal (10 μM); resting [Ca2+]cyto = 97.3 ± 8.3% basal (10 μM)). As unfavourable cardiovascular outcomes, possibly related to disrupted cellular Ca2+ homeostasis, have been linked to Rosiglitazone, this effect may be of clinical interest. In contrast, in PPRE-luciferase reporter-gene assays, Rosiglitazone induced non-dose-dependent PPARγ-dependent effects (1 μM: 152.5 ± 4.9% basal; 10 μM: 136.1 ± 5.1% basal (p < 0.05 for 1 μM vs. 10 μM)). Thus, we conclude that Rosiglitazone can exert PPARγ-independent non-genomic effects, such as the SERCA2b inhibition seen here, but that long-term Rosiglitazone treatment did not perturb resting [Ca]cyto in this study.  相似文献   

6.
Parrotia persica C.A. Meyer (Persian ironwood) is a deciduous tree of the family Hamamelidaceae, native to northern Iran and endemic to the Alborz Mountains. The study objectives were to assess the current status and distribution of Persian ironwood by characterizing four forest stands where the tree was either a dominant or co-dominant species. Species richness within the stands varied from 3 to 16 woody species and from 9 to 27 understory species. Basal area varied between 37 m2/ha and 77 m2/ha and tree density varied from 320 to 367 stems/ha. Parrotia persica represented 63-86% of the relative dominance and 41-100% of the relative density. In non-pure P. persica stands, other important tree species include Fagus orientalis and Carpinus betulus. Parrotia persica regenerates mainly by sprouts and coppicing. Conservation of relict forests, such as the Persian ironwood forests of the Alborz Mountains, is of particular concern because they represent the only natural occurrence of this species in the world. Anthropogenic disturbance, in the form of timber harvesting, livestock grazing, and clearing forest land for agriculture appear to be the largest threats to Parrotia persica's future.  相似文献   

7.
We present an overview of long-term changes in the floristic composition and growth areas in L. Peipsi (3555 km2, unregulated water level) that have occurred since the 1960s and a list of plant taxa containing 140 species of higher plants and 4 charophytes. A significant correlation was found between the relative abundance and frequency per stations (Fs) (Rs = 0.93). Data on five inhabitants of the eulittoral revealed significant (p < 0.05) inter-annual differences in Fs. Comparison of the data of Fs for 67 taxa for 1970-1980 (87 stations) and 1997-2007 (139 stations) showed a significant change in the Fs distribution (p < 0.03) and a decline (p < 0.05) for 20 taxa; for 15 species Fs had decreased two times or more. However, 14 of the markedly declined taxa, e.g. the long-term dominating submergent Potamogeton perfoliatus, belong still among the top 33 in the list. A significant (χ2 = 11.8; p < 0.028) change was observed in the species number of different frequency classes. The number of taxa in the Fs class 46-100 (92)% was 17 in 1970-1980 but only 3 in 1997-2007. The top of the list of macrophytes is dominated by circumpolar species and vicariants. Impoverishment of the flora in the course of eutrophication is expressed by the decrease in Fs; at the same time, the total number of species had not changed. Among the 20 declined taxa 14 are characteristic of the temporarily flooded and/or shallow-water zone of eutrophic water bodies (amphibious and emergent plants); the remaining taxa are shallow-water submergents. The simpliest explanation for their decrease is the expansion of thick reeds occupying suitable eulittoral habitats.  相似文献   

8.

Background

CXCL10 may contribute to the host immune response against the hepatitis C virus (HCV), liver disease progression, and response to HCV antiviral therapy. The aim of our study was to analyze the relationship among virological, immunological, and clinical characteristics with plasma CXCL10 levels in human immunodeficiency virus (HIV)/HCV-coinfected patients.

Methods

We carried out a cross-sectional study on 144 patients. CXCL10 and insulin were measured using an immunoassay kit. The degree of insulin resistance was estimated for each patient using the homeostatic model assessment (HOMA) method. Insulin resistance was defined as a HOMA index higher than or equal to 3.8. Aspartate aminotransferase (AST) to platelet ratio (APRI), FIB-4, Forns index, HGM1, and HGM2 were calculated.

Results

The variables associated with log10 CXCL10 levels by univariate analysis were age (b = 0.013; p = 0.023), prior AIDS-defining condition (b = 0.127; p = 0.045), detectable plasma HIV viral load (b = 0.092; p = 0.006), log10 HOMA (b = 0.216; p = 0.002), HCV-genotype 1 (b = 0.114; p = 0.071), and liver fibrosis assessed by all non-invasive indexes (log10 APRI (b = 0.296; p = 0.001), log10 FIB-4 (b = 0.436; p < 0.001), log10 Forns index (b = 0.591; p < 0.001), log10 HGM1 (b = 0.351; p = 0.021), and log10 HGM2 (b = 0.215; p = 0.018)). However, in multivariate analysis, CXCL10 levels were only associated with HOMA, detectable plasma HIV viral load, HCV-genotype 1 and FIB-4 (R-square = 0.235; p < 0.001).

Conclusion

Plasma CXCL10 levels were influenced by several characteristics of patients related to HIV and HCV infections, insulin resistance, and liver fibrosis, indicating that CXCL10 may play an important role in the pathogenesis of both HCV and HIV infections.  相似文献   

9.
Knowledge of the social structure of a population is important for a range of fundamental and applied purposes. Group characteristics and population structure of chital (Axis axis) and sambar (Rusa unicolor) were studied in a deciduous habitat of Mudumalai Tiger Reserve, Western Ghats, India, during 2008-2009. Vehicle transects were monitored monthly to gather information on group-size and age-sex composition of chital and sambar. The average mean group size and crowding for chital and sambar was 13.1 ± 0.5 (n = 1020), 3.6 ± 0.2 (n = 377) and 33.3, 11.0 respectively. The average adult male:adult female:fawn ratio was 63.4:100:22.3 (n = 9391) and 43.9:100:23.7 (n = 1023) in chital and sambar respectively. The mean group size of chital and sambar varied significantly between seasons (Kruskal-Wallis test, p = <0.001). Peak fawning season was observed from February to May for chital and May to August for sambar. Group's sex-age composition influenced group formation in both species between seasons at different level. Adult male and fawn were the important predicting variables of change in group size. Skewed female sex ratio was probably due to selective male predation by large predators. Although fawning occurred throughout the year, both species showed seasonality in fawning. The above mentioned patterns differed between species depending upon their ecological adaptation in foraging strategies and habitat preference.  相似文献   

10.
Substitution reaction of chloro η6-arene ruthenium N∩O-base complexes [(η6-arene)Ru(N∩O)Cl] [N∩O = pyrazine-2-carboxylic acid (pca-H), 8-hydroxyquinoline (hq-H); arene = p-iPrC6H4Me, N∩O = hq (1); arene = C6Me6, N∩O = hq (2)] with NaN3 yield the neutral arene ruthenium azido complexes of the general formula [(η6-arene)Ru(N∩O)N3] [N∩O = pca, arene = p-iPrC6H4Me (3), arene = C6Me6 (4); N∩O = hq, arene = p-iPrC6H4Me (5), arene = C6Me6 (6)]. These complexes undergo [3 + 2] dipolar cycloaddition reaction with activated alkynes dimethyl and diethyl acetylenedicarboxylates to yield the arene triazole complexes [(η6-arene)Ru(N∩O){N3C2(CO2R)2}] [N∩O = pca, R = Me, arene = p-iPrC6H4Me (7), C6Me6 (8); R = Et, arene = p-iPrC6H4Me (9), C6Me6 (10); N∩O = hq, R = Me, arene = p-iPrC6H4Me (11) C6Me6 (12); R = Et, arene = p-iPrC6H4Me (13), C6Me6 (14)]. On the bases of proton NMR study, in the above triazole complexes N(2) isomers are assigned with dimethylacetylenedicarboxylate whereas N(1) isomers with diethylacetylenedicarboxylate. All complexes have been characterized by IR and NMR spectroscopy as well as by elemental analysis. The molecular structures of the azido complexes [(η6-p-iPrC6H4Me)Ru(pca)N3] (3), [(η6-p-iPrC6H4Me)Ru(hq)N3] (5) and [(η6-C6Me6)Ru(hq)N3] (6) have been established by single crystal X-ray diffraction studies.  相似文献   

11.
Leaves of 26 grass, herb, shrub and tree species were collected from mesotrophic grasslands to assess natural variability in bulk, fatty acid and monosaccharide δ13C values under different grazing management (cattle- or deer-grazed) on three sample dates (May, July and October) such that interspecific and spatiotemporal variations in whole leaf tissues and compound-specific δ13C values could be determined. The total mean leaf bulk δ13C value for plants was −28.9‰ with a range of values spanning 7.5‰. Significant interspecific variation between bulk leaf δ13C values was only determined in October (P = <0.001) when δ13C values of the leaf tissues from both sites was on average 1.5‰ depleted compared to during July and May. Samples from May were significantly different between fields (P = 0.03) indicating an effect from deer- or cattle-grazing in young leaves. The average individual monosaccharide δ13C value was 0.8‰ higher compared with whole leaf tissues. Monosaccharides were the most abundant components of leaf biomass, i.e. arabinose, xylose, mannose, galactose and glucose, and therefore, fluctuations in their individual δ13C values had a major influence on bulk δ13C values. An average depletion of ca. 1‰ in the bulk δ13C values of leaves from the deer-grazed field compared to the cattle-grazed field could be explained by a general depletion of 1.1‰ in glucose δ13C values, as glucose constituted >50% total leaf monosaccharides. In October, δ13C values of all monosaccharides varied between species, with significant variation in δ13C values of mannose and glucose in July, and mannose in May. This provided an explanation for the noted variability in the tissue bulk δ13C values observed in October 1999. The fatty acids C16:0, C18:2 and C18:3 were highly abundant in all plant species. Fatty acid δ13C values were lower than those of bulk leaf tissues; average values of −37.4‰ (C16:0), −37.0‰ (C18:2) and −36.5‰ (C18:3) were determined. There was significant interspecific variation in the δ13C values of all individual fatty acids during October and July, but only for C18:2 in May (P = <0.05). This indicated that seasonal trends observed in the δ13C values of individual fatty acids were inherited from the isotopic composition of primary photosynthate. However, although wide diversity in δ13C values of grassland plants ascribed to grazing management, interspecific and spatiotemporal influences was revealed, significant trends (P = <0.0001) for fatty acid and monosaccharide δ13C values: δ13C16:0 < δ13C18:2 < δ13C18:3 and δ13Carabinose > δ13Cxylose > δ13Cglucose > δ13Cgalactose, respectively, previously described, appear consistent across a wide range of species at different times of the year in fields under different grazing regimes.  相似文献   

12.
Light use efficiency (LUE) is an important variable in carbon cycle and climate change research. We present an investigation of remotely estimating midday LUE using the green chlorophyll index (CIgreen) derived from the cloud-free Moderate Resolution Imaging Spectroradiometer (MODIS) images in maize, coniferous forest and grassland. Similar temporal patterns are observed in both canopy chlorophyll content and midday LUE which indicates that the chlorophyll content in the maize canopy servers as a proxy of midday LUE (R2 = 0.736, p < 0.001). Therefore, the CIgreen, tested as a good indicator of canopy chlorophyll content (R2 = 0.840, p < 0.001), has been demonstrated to be a reliable candidate in providing reasonable estimates of midday LUE with determination coefficient R2 equals to 0.820 and a root mean square error (RMSE) of 0.002 mol CO2 per mol incident photosynthetic photon flux density (PPFD). Further validation of the prediction model derived from the maize site demonstrates that the CIgreen has potential to be applied in the coniferous forest and grassland ecosystems with RMSE of 0.005 and 0.004 mol CO2 mol−1 PPFD, respectively. A comparison analysis between different vegetation types is included and these results could be helpful in the development of future LUE and terrestrial models.  相似文献   

13.
Binding of the utmost N-terminus of essential myosin light chains (ELC) to actin slows down myosin motor function. In this study, we investigated the binding constants of two different human cardiac ELC isoforms with actin. We employed circular dichroism (CD) and surface plasmon resonance (SPR) spectroscopy to determine structural properties and protein–protein interaction of recombinant human atrial and ventricular ELC (hALC-1 and hVLC-1, respectively) with α-actin as well as α-actin with alanin-mutated ELC binding site (α-actinala3) as control. CD spectroscopy showed similar secondary structure of both hALC-1 and hVLC-1 with high degree of α-helicity. SPR spectroscopy revealed that the affinity of hALC-1 to α-actin (KD = 575 nM) was significantly (p < 0.01) lower compared with the affinity of hVLC-1 to α-actin (KD = 186 nM). The reduced affinity of hALC-1 to α-actin was mainly due to a significantly (p < 0.01) lower association rate (kon: 1018 M−1 s−1) compared with kon of the hVLC-1/α-actin complex interaction (2908 M−1 s−1). Hence, differential expression of ELC isoforms could modulate muscle contractile activity via distinct α-actin interactions.  相似文献   

14.
The ruthenium(II) hexaaqua complex [Ru(H2O)6]2+ reacts with dihydrogen under pressure to give the η2-dihydrogen ruthenium(II) pentaaqua complex [Ru(H2)(H2O)5]2+.The complex was characterized by 1H, 2H and 17O NMR: δH = −7.65 ppm, JHD = 31.2 Hz, δO = −80.4 ppm (trans to H2) and δO = −177.4 ppm (cis to H2).The H-H distance in coordinated dihydrogen was estimated to 0.889 Å from JHD, which is close to the value obtained from DFT calculations (0.940 Å).Kinetic studies were performed by 1H and 2H NMR as well as by UV-Vis spectroscopy, yielding the complex formation rate and equilibrium constants: kf = (1.7 ± 0.2) × 10−3 kg mol−1 s−1 and Keq = 4.0 ± 0.5 mol kg−1.The complex formation rate with dihydrogen is close to values reported for other ligands and thus it is assumed that the reaction with dihydrogen follows the same mechanisn (Id).In deuterated water, one can observe that [Ru(H2)(H2O)5]2+ catalyses the hydrogen exchange between the solvent and the dissolved dihydrogen.A hydride is proposed as the intermediate for this exchange.Using isotope labeling, the rate constant for the hydrogen exchange on the η2-dihydrogen ligand was determined as k1 = (0.24 ± 0.04) × 10−3 s−1.The upper and lower limits of the pKa of the coordinated dihydrogen ligand have been estimated:3 < pKa < 14.  相似文献   

15.
Sadaruddin Biswas 《HOMO》2010,61(4):271-276
One of the greatest problems facing developing countries, including rural India, is undernutrition in terms of stunting among under 5-year-old children. However, there exists scanty information on the prevalence of stunting among preschool children in India and in particular in West Bengal. This study investigated prevalence of stunting and identified the predictor(s) of stunting among 1-5-year-old Bengalee rural preschool children of Integrated Child Development Services (ICDS) centres. This cross-sectional study was undertaken at different ICDS centres of Chapra Block, Nadia District, West Bengal, India. A total of 673 preschool children (323 boys and 350 girls), aged 1-5 years were selected from 30 randomly selected ICDS centres to study the impact of parents’ educational status and child birth order on stunting. The overall (age and sex combined) rate of stunting was 39.2%. Child birth order (BO) (χ2 = 14.10, df = 1, p < 0.001), father educational status (FES) (χ2 = 21.11, p < 0.001) and mother educational status (MES) (χ2 = 14.34, df = 1, p > 0.001) were significantly associated with the prevalence of stunting among girls. Logistic regression analyses revealed that both FES (Wald = 19.97, p < 0.001) as well as MES (Wald = 13.95, p < 0.001) were strong predictors of stunting among girls. Similarly BO (Wald = 13.71, p < 0.001) was a strong predictor of stunting among girls. Girls with ≥3rd BO had significantly higher risk (OR = 2.49, CI = 1.54-4.03) of stunting than those with ≤2nd BO. Moreover, girls with FES lower than secondary level had significantly (OR = 3.30, CI = 1.96-5.58) higher rate of stunting than those with FES ≥ secondary level. Similarly, girls with MES < secondary level had significantly (OR = 2.50, CI = 1.54-4.03) higher rate of stunting than those with FES ≥ secondary level.In conclusion our study revealed that BO as well as parents’ educational status were strong predictors of stunting among girls but not boys. Sex discrimination could be a likely cause for this sex difference in the impact of BO and parents’ educational status.  相似文献   

16.
Industrial wastewater treatment comprises several processes to fulfill the discharge permits or to enable the reuse of wastewater. For tannery wastewater, constructed wetlands (CWs) may be an interesting treatment option. Two-stage series of horizontal subsurface flow CWs with Phragmites australis (UP series) and Typha latifolia (UT series) provided high removal of organics from tannery wastewater, up to 88% of biochemical oxygen demand (BOD5) (from an inlet of 420 to 1000 mg L−1) and 92% of chemical oxygen demand (COD) (from an inlet of 808 to 2449 mg L−1), and of other contaminants, such as nitrogen, operating at hydraulic retention times of 2, 5 and 7 days. No significant (P < 0.05) differences in performance were found between both the series. Overall mass removals of up to 1294 kg COD ha−1 d−1 and 529 kg BOD5 ha−1 d−1 were achieved for a loading ranging from 242 to 1925 kg COD ha−1 d−1 and from 126 to 900 kg BOD5 ha−1 d−1. Plants were resilient to the conditions imposed, however P. australis exceeded T. latifolia in terms of propagation.  相似文献   

17.
Oxygen respiration rates of benthic foraminifera are still badly known, mainly because they are difficult to measure. Oxygen respiration rates of seventeen species of benthic foraminifera were measured using microelectrodes and calculated on the basis of the oxygen fluxes measured in the vicinity of the foraminiferal specimens. The results show a wide range of oxygen respiration rates for the different species (from 0.09 to 5.27 nl cell−1 h−1) and a clear correlation with foraminiferal biovolume showed by the power law relationship: R = 3.98 10−3 BioVol0.88 where the oxygen respiration rate (R) is expressed in nl O2 h−1 and in μm3 biovolume (BioVol) (n = 44, R2 = 0.72, F = 114, p < 0.0001). The results expressed per biovolume unit (1.82 to 15.7 nl O2 10−8 μm−3 h−1) allow us to compare our data with the previous published data showing similar ranges. A comparison with available data for other microbenthos groups (nematodes, copepods, ostracods, ciliates and flagellates) suggests that benthic foraminifera have a lower oxygen respiration rates per unit biovolume. The total contribution of benthic foraminifera to the aerobic mineralisation of organic matter is estimated for the studied areas. The results suggest that benthic foraminifera play only a minor role (0.5 to 2.5%) in continental shelf environments, which strongly contrasts with their strong contribution to anaerobic organic matter mineralisation, by denitrification, in the same areas.  相似文献   

18.
From 2005 to 2007, we established bird-proof enclosures in a small, shallow and semi-permanent lake, lacking fish, at Brown Moss, Shropshire, UK, to investigate the effects of aquatic birds on seasonal growth of submerged and emergent macrophytes. The highest density of birds on the lake was in winter (110 individuals ha−1) and the lowest in summer 2005 (6 ha−1). Plant growth varied with season but there were significantly different (F = 8.03, p < 0.05, df = 1) standing crops of macrophytes between bird-proof enclosures (proportion of volume occupied, 0.47 ± 0.04) and control treatments (0.36 ± 0.11). Different densities of birds occurred in different areas and this was reflected in their effects. Ducks, mainly mallard (Anas platyrhynchos, Linnaeus), and teal (Anas crecca, Linnaeus), damaged plants by direct consumption, uprooting and trampling, whereas larger birds, such as mute swan (Cygnus olor, Gmelin), were able to remove Typha latifolia (Linnaeus). In summer, grazing pressure was reduced as the population of birds declined. Waterfowl caused seasonal impacts on the re-development of the water plant community. However, waterfowl herbivory had low potential to shift a macrophyte-dominated state into a phytoplankton-dominated state because aquatic plants could recover, during the growing season, when bird populations declined.  相似文献   

19.
Rate and equilibrium constants at 25 °C, pH ∼ 1, and ionic strength 0.10 for hydrolysis of the two non-equivalent chlorides of dichloro[S-methyl-l-cysteine(N,S)]platinum(II) isomers, denoted [PtCl2(SmecysH)], and the resultant chloro-aqua species have been determined by NMR, potentiometric, and spectrophotometric methods. Though hydrolysis constants, Kh, for the two chlorides are similar (pKh = 4-5), the rate of hydrolysis of the chloride trans to coordinated S, kh = 3.4 × 10−3 s−1, is 2-3 orders of magnitude faster than the kh for the other chloride, 2.3 × 10−6 s−1, and for the cancer drug cisplatin, cis-[PtCl2(NH3)2], 5.2 × 10−5 s−1. Relative rates of hydrolysis determined under three different experimental conditions (pH ∼ 1 in 0.10 M HNO3, high pH in 0.10 M NaOH, and at low pH with Ag+ assistance) are consistent: the Cl trans to S is 100-1000 times more labile than the Cl cis to S. Potentiometric and NMR methods were also used to estimate pKa values of all aqua species, which are comparable to values reported for corresponding aqua species derived from cisplatin.  相似文献   

20.
The flowers of 23 species of grass and herb plants were collected from a mesotrophic grassland to assess natural variability in bulk, monosaccharide and fatty acid δ13C values from one plant community and were compared with previous analyses of leaves from the same species. The total mean bulk δ13C value of flower tissues was −28.1‰, and there was no significant difference between the mean δ13Cflower values for grass (−27.8‰) and herb (−28.2‰) species. On average bulk δ13Cflower values were 1.1‰ higher than bulk δ13Cleaf values, however, the δ13Cflower and δ13Cleaf values of grasses did not differ between organs suggesting that carbon isotope discrimination is different in grass and herb species. The abundance of different monosaccharides abundance varied between plant types, i.e. xylose concentrations in the grass flowers were as high as 40%, compared with up to 15% in the herb species, but the general relationship δ13Carabinose > δ13Cxylose > δ13Cglucose > δ13Cgalactose which had been observed in leaves was similar in flowers (total mean δ13C values = −25.9‰, −27.2‰, −28.8‰ and −28.1‰, respectively). However, the average 5.4‰ depletion in the δ13C values of the C16:0, C18:2 and C18:3 fatty acids in flowers compared to bulk tissue was significantly greater than observed for leaves. The trend C16:0 < C18:2 < C18:3 previously observed in leaves was also observed in grass flowers (δ13CC16:0 = −33.8‰; δ13CC18:2 = −33.1‰; δ13CC18:3 = −34.2‰) but not herb flowers (δ13CC16:0 = −34.1‰; δ13CC18:2 = −32.4‰; δ13CC18:3 = −34.5‰). We conclude: (i) that the biological processes influencing carbon isotope discrimination in grass flowers are different from herbs flowers; and, (ii) that a range of post-photosynthetic fractionation effects caused the observed differences between flower and leaf δ13C values, especially the significant 13C-depletion in flower fatty acid δ13C values.  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