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.
A procedure to measure exchange rates of fast exchanging protein amide hydrogens by time-resolved NMR spectroscopy following in situ initiation of the reaction by diluting a native protein solution into an exchanging deuterated buffer is described. The method has been used to measure exchange rates of a small set of amide hydrogens of reduced cytochrome c, maintained in a strictly anaerobic atmosphere, in the presence of an otherwise inaccessible range of guanidinium deuterochloride concentrations. The results for the measured protons indicate that hydrogen exchange in the unfolding transition region of cytochrome c reach the EX2 limit, but emphasize the difficulty in interpretation of the exchange mechanism in protein hydrogen exchange studies. Comparison of free energies of structure opening for the measured hydrogens with the global unfolding free energy monitored by far-UV CD measurements has indicated the presence of at least one partially unfolded equilibrium species of reduced cytochrome c. The results provide the first report of measurement of free energy of opening of structure to exchange in the 0–2-kcal/mol range. Proteins 32:241–247, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

2.
3.
Amide hydrogen/deuterium exchange rates were measured as a function of pH and urea for 37 slowly exchanging amides in the beta-trefoil protein hisactophilin. The rank order of exchange rates is generally maintained under different solution conditions, and trends in the pH and urea dependence of exchange rates are correlated with the rank order of exchange rates. The observed trends are consistent with the expected behavior for exchange of different amides via global and/or local unfolding. Analysis of the pH dependence of exchange in terms of rate constants for structural opening and closing reveals a wide range of rates in different parts of the hisactophilin structure. The slowest exchanging amides have the slowest opening and closing rates. Many of the slowest exchanging amides are located in trefoil 2, but there are also some slow exchanging amides in trefoils 1 and 3. Slow exchangers tend to be near the interface between the beta-barrel and the beta-hairpin triplet portions of this single-domain structure. The pattern of exchange behaviour in hisactophilin is similar to that observed previously in interleukin-1 beta, indicating that exchange properties may be conserved among beta-trefoil proteins. Comparisons of opening and closing rates in hisactophilin with rates obtained for other proteins reveal clear trends for opening rates; however, trends in closing rates are less apparent, perhaps due to inaccuracies in the values used for intrinsic exchange rates in the data fitting. On the basis of the pH and urea dependence of exchange rates and optical measurements of stability and folding, EX2 is the main exchange mechanism in hisactophilin, but there is also evidence for varying levels of EX1 exchange at low and high pH and high urea concentrations. Equilibrium intermediates in which subglobal portions of structure are cooperatively disrupted are not apparent from analysis of the urea dependence of exchange rates. There is, however, a strong correlation between the Gibbs free energy of opening and the denaturant dependence of opening for all amides, which suggests exchange from a continuum of states with different levels of structure. Intermediates are not very prominent either in equilibrium exchange experiments or in quenched-flow kinetic studies; hence, hisactophilin may not form partially folded states as readily as IL-1 beta and other beta-trefoil proteins.  相似文献   

4.
Gas/water interfaces (such as air bubbles or foam) are detrimental to the stability of proteins, often causing aggregation. This represents a potential problem for industrial processes, for example, the production and handling of protein drugs. Proteins possess surfactant-like properties, resulting in a high affinity for gas/water interfaces. The tendency of previously buried nonpolar residues to maximize contact with the gas phase can cause significant structural distortion. Most earlier studies in this area employed spectroscopic tools that could only provide limited information. Here we use hydrogen/deuterium exchange (HDX) mass spectrometry (MS) for probing the conformational dynamics of the model protein myoglobin (Mb) in the presence of N2 bubbles. HDX/MS relies on the principle that unfolded and/or highly dynamic regions undergo faster deuteration than tightly folded segments. In bubble-free solution Mb displays EX2 behavior, reflecting the occurrence of short-lived excursions to partially unfolded conformers. A dramatically different behavior is seen in the presence of N2 bubbles; EX2 dynamics still take place, but in addition the protein shows EX1 behavior. The latter results from interconversion of the native state with conformers that are globally unfolded and long-lived. These unfolded species likely correspond to Mb that is adsorbed to the surface of gas bubbles. N2 sparging also induces aggregation. To explain the observed behavior we propose a simple model, that is, “semi-unfolded” ↔ “native” ↔ “globally unfolded” → “aggregated”. This model quantitatively reproduces the experimentally observed kinetics. To the best of our knowledge, the current study marks the first exploration of surface denaturation phenomena by HDX/MS.  相似文献   

5.
A DNA genomic library constructed from Bacillus stearothermophilus, a gram-positive, facultative thermophilic aerobe that secretes a thermostable β-mannanase, was screened for mannan hydrolytic activity. Recombinant β-mannanase activity was detected on the basis of the clearing of halos around Escherichia coli colonies grown on a dye-labelled substrate, Remazol brilliant blue-locust bean gum. The nucleotide sequence of the mannanase gene, manF, corresponded to an open reading frame of 2,085 bp that codes for a 32-amino-acid signal peptide and a mature protein with a molecular mass of 76,089 Da. From sequence analysis, ManF belongs to glycosyl hydrolase family 5 and exhibits higher similarity to eukaryotic than to bacterial mannanases. The manF coding sequence was subcloned into the pH6EX3 expression plasmid and expressed in E. coli as a recombinant fusion protein containing a hexahistidine N-terminal sequence. The fusion protein has thermostability similar to the native enzyme and was purified by Ni2+ affinity chromatography. The values for the kinetic parameters Vmax and Km were 384 U/mg and 2.4 mg/ml, respectively, for the recombinant mannanase and were comparable to those of the native enzyme.  相似文献   

6.
Net photosynthetic CO2 uptake (P N ) and transpiration (E) rates through adaxial and abaxial leaf surfaces were measured gasometrically on the area of maize leaf blades of different insertion levels in the phase of male florescence and in the phase of spadices development. The distribution pattern of gas exchange on the area of all measured leaf blades found in the phase of male florescence (significantly lower value in the basal parts) was changed in the phase of spadices development: PN andE of abaxial surfaces of older leaves,i.e. 9th, 10th (under which the spadices were growing) and 11th, rapidly increased in the basal parts in comparison with the middle ones and positively influenced the gas exchange of the whole leaves. It appears that PN may be maintained or increased in the older leaves in the phase of spadices development to prevent shortage of photosynthates.  相似文献   

7.
Styrax caporum is a native shrub from the Brazilian savanna. Most of its leaves are diaheliotropic, whereas some are paraheliotropic, mainly at noon. A previous study of this species revealed higher stomatal conductance (gs) and transpiration rates (E) in para- compared to diaheliotropic leaves, and a rise in CO2 assimilation rates (A) with an increase of irradiance for paraheliotropic leaves. We hypothesized that this species exploits the paraheliotropism to enhance the light use efficiency, and that it is detected only if gas exchange is measured with light interception by both leaf surfaces. Gas exchange was measured with devices that enabled light interception on only one of the leaf surfaces and with devices that enabled light interception by both leaf surfaces. Water relations, relative reflected light intensity, leaf temperature (Tl), and leaf anatomical analyses were also performed. When both leaf surfaces were illuminated, a higher A, E, and gs were observed in para- compared to diaheliotropic leaves; however, A did not depend on gs, which did not influence CO2 accumulation in the stomatal cavity (Ci). When only the adaxial leaf surface was illuminated, a greater A was detected for para- than for diaheliotropic leaves only at 11:00 h; no differences in Tl were observed between leaf types. Light curves revealed that under non-saturating light the adaxial side of paraheliotropic leaves had higher A than the abaxial side, but they showed similar values under saturating light. Although the abaxial leaf side was highly reflective, both surfaces presented the same response pattern for green light reflection, which can be explained by the compact spongy parenchyma observed in the leaves, increasing light use efficiency in terms of CO2 consumption for paraheliotropic leaves. We propose that paraheliotropism in S. camporum is not related to leaf heat avoidance or photoprotection.  相似文献   

8.
Escherichia coli CspA is a small all-beta-sheet protein that folds fast (tau = 4 ms) via an apparent two-state mechanism. Our previous studies have shown that a large aromatic cluster on the surface of the protein participates in the rate-limiting step of folding and thus may be part of the folding nucleus of this protein. To obtain a more detailed picture of molecular events at the peptide backbone during unfolding and folding of CspA, we used native state hydrogen exchange and nuclear magnetic resonance spectroscopy (NMR). The experiments with native CspA were performed over a range of pH values from low pH, where exchange is governed by a rapid equilibrium before chemical exchange (EX2 exchange), to high pH, where exchange is dictated by the rate of unfolding (EX1 exchange). Rates of folding and unfolding were determined for 11 residues. The distribution of rates of folding within the structure of CspA suggests that hairpin turns, including one near the aromatic cluster, may nucleate the folding of CspA.  相似文献   

9.
To search for submolecular foldon units, the spontaneous reversible unfolding and refolding of staphylococcal nuclease under native conditions was studied by a kinetic native-state hydrogen exchange (HX) method. As for other proteins, it appears that staphylococcal nuclease is designed as an assembly of well-integrated foldon units that may define steps in its folding pathway and may regulate some other functional properties. The HX results identify 34 amide hydrogens that exchange with solvent hydrogens under native conditions by way of large transient unfolding reactions. The HX data for each hydrogen measure the equilibrium stability (ΔGHX) and the kinetic unfolding and refolding rates (kop and kcl) of the unfolding reaction that exposes it to exchange. These parameters separate the 34 identified residues into three distinct HX groupings. Two correspond to clearly defined structural units in the native protein, termed the blue and red foldons. The remaining HX grouping contains residues, not well separated by their HX parameters alone, that represent two other distinct structural units in the native protein, termed the green and yellow foldons. Among these four sets, a last unfolding foldon (blue) unfolds with a rate constant of 6 × 10− 6 s− 1 and free energy equal to the protein's global stability (10.0 kcal/mol). It represents part of the β-barrel, including mutually H-bonding residues in the β4 and β5 strands, a part of the β3 strand that H-bonds to β5, and residues at the N-terminus of the α2 helix that is capped by β5. A second foldon (green), which unfolds and refolds more rapidly and at slightly lower free energy, includes residues that define the rest of the native α2 helix and its C-terminal cap. A third foldon (yellow) defines the mutually H-bonded β1-β2-β3 meander, completing the native β-barrel, plus an adjacent part of the α1 helix. A final foldon (red) includes residues on remaining segments that are distant in sequence but nearly adjacent in the native protein. Although the structure of the partially unfolded forms closely mimics the native organization, four residues indicate the presence of some nonnative misfolding interactions. Because the unfolding parameters of many other residues are not determined, it seems likely that the concerted foldon units are more extensive than is shown by the 34 residues actually observed.  相似文献   

10.
Tritium exchange was used as a probe of transfer RNA structure in experiments with unfractionated tRNA (tRNAUnfrac and homogeneous tRNA3Leu from bakers' yeast. Exchange kinetics were measured over a range of ionic conditions that vary in ability to stabilize the secondary and tertiary structure of tRNA. The native conformations of both samples show the same kinetics of exchange. The kinetics for tRNA3Leu trapped in a denatured state in a “native” solvent are much faster, reflecting the conformation and not the ionic medium. In 0.1 M-Na+, where tRNA3Leu is denatured, the kinetics for tRNAUnfrac are intermediate between those for native and denatured tRNA3Leu, suggesting that in this solvent at 0 °C some tRNAs are denatured whereas other are still native. Upon further lowering of Na+ concentration, tRNAUnfrac shows increasingly faster exchange, suggesting complete electrostatic denaturation of the tertiary structure of all the tRNAs in the sample, and even disruption of secondary structure.Extrapolation of the essentially linear early-time kinetics to zero time provides minimal estimates of the number of slowly exchanging hydrogens. For native tRNA3Leu the number is 111±2 hydrogens, whereas for the trapped denatured conformation it is only 95±2. This difference reflects a smaller number of hydrogen-bonded bases in the denatured conformation. In 1 M-Na+, 101±2 slowly exchanging hydrogens are found for the native tRNA3Leu conformation, suggesting an incompletely formed native structure. For native tRNAUnfrac the comparable number is 101±3. These numbers of slowly exchanging hydrogens in the native conformations are consistent with tertiary structural hydrogen-bonding. Furthermore, this tertiary structure must be responsible for the slower exchange by native tRNA. The observed numbers of exchangeable hydrogens provide a basis for comparison of hydrogen-bonding interactions in native and denatured tRNA conformations.The mechanism of renaturation was also investigated, using tritium exchange as a monitor of perturbation of base pairing during the transition. When tRNAUnfrac in low Na+ is renatured by addition of Mg2+ during tritium exchangeout, a burst of exchange or “spillage” of tritium is detected. This suggests that a fraction of the base pairs of the rapidly renaturing tRNAs in the mixture is disrupted during renaturation. In that event, and by analogy with tRNA3Leu, part of the base-pairing arrangement of the denatured conformations may not be preserved in the native state; and if the native conformation includes the full “cloverleaf” pattern of secondary structure, that pattern may not be intact in some denatured conformations.  相似文献   

11.

Background

The superficial vein of the resting limb constricts sympathetically during exercise. Central command is the one of the neural mechanisms that controls the cardiovascular response to exercise. However, it is not clear whether central command contributes to venous vessel response during exercise. Tendon vibration during static elbow flexion causes primary muscle spindle afferents, such that a lower central command is required to achieve a given force without altering muscle force. The purpose of this study was therefore to investigate whether a reduction in central command during static exercise with tendon vibration influences the superficial venous vessel response in the resting limb.

Methods

Eleven subjects performed static elbow flexion at 35% of maximal voluntary contraction with (EX + VIB) and without (EX) vibration of the biceps brachii tendon. The heart rate, mean arterial pressure, and rating of perceived exertion (RPE) in overall and exercising muscle were measured. The cross-sectional area (CSAvein) and blood velocity of the basilic vein in the resting upper arm were assessed by ultrasound, and blood flow (BFvein) was calculated using both variables.

Results

Muscle tension during exercise was similar between EX and EX + VIB. However, RPEs at EX + VIB were lower than those at EX (P <0.05). Increases in heart rate and mean arterial pressure during exercise at EX + VIB were also lower than those at EX (P <0.05). CSAvein in the resting limb at EX decreased during exercise from baseline (P <0.05), but CSAvein at EX + VIB did not change during exercise. CSAvein during exercise at EX was smaller than that at EX + VIB (P <0.05). However, BFvein did not change during the protocol under either condition. The decreases in circulatory response and RPEs during EX + VIB, despite identical muscle tension, showed that activation of central command was less during EX + VIB than during EX. Abolishment of the decrease in CSAvein during exercise at EX + VIB may thus have been caused by a lower level of central command at EX + VIB rather than EX.

Conclusion

Diminished central command induced by tendon vibration may attenuate the superficial venous vessel response of the resting limb during sustained static arm exercise.  相似文献   

12.
The dormant cysts of Artemia undergo cycles of hydration-dehydration without losing viability. Therefore, Artemia cysts serve as an excellent intact cellular system for studying the dynamics of water-protein interactions as a function of hydration. Deuterium spin-lattice (T1) and spin-spin (T2) relaxation times of water in cysts hydrated with D2O have been measured for hydrations between 1.5 and 0.1 g of D2O per gram of dry solids. When the relaxation rates (I/T1, I/T2) of 2H and 17O are plotted as a function of the reciprocal of hydration (1/H), an abrupt change in slope is observed near 0.6 g of D2O (or H2 17O)/gram of dry solids, the hydration at which conventional metabolism is activated in this system. The results have been discussed in terms of the two-site and multisite exchange models for the water-protein interaction as well as protein dynamics models. The 2H and 17O relaxation rates as a function of hydration show striking similarities to those observed for anisotropic motion of water molecules in protein crystals.

It is suggested here that although the simple two-site exchange model or n-site exchange model could be used to explain our data at high hydration levels, such models are not adequate at low hydration levels (<0.6 g H2O/g) where several complex interactions between water and proteins play a predominant role in the relaxation of water nuclei. We further suggest that the abrupt change in the slope of I/T1 as a function of hydration in the vicinity of 0.6 g H2O/g is due to a change in water-protein interactions resulting from a variation in the dynamics of protein motion.

  相似文献   

13.
Isotopic exchange kinetics at equilibrium for E. coli native aspartate transcarbamylase at pH 7.8, 30 °C, are consistent with an ordered BiBi substrate binding mechanism. Carbamyl phosphate binds before l-Asp, and carbamyl-aspartate is released before inorganic phosphate. The rate of [14C]Asp C-Asp exchange is much faster than [32P]carbamyl phosphate Pi exchange. Phosphate, and perhaps carbamyl phosphate, appears to bind at a separate modifier site and prevent dissociation of active-site bound Pi or carbamyl phosphate. Initial velocity studies in the range of 0–40 °C reveal a biphasic Arrhenius plot for native enzyme: Ea (>15 °C) = 6.3 kcal/ mole and Ea (<15 °C) = 22.1 kcal/mole. Catalytic subunits show a monophasic plot with Ea ? 20.2 kcal/mole. This, with other data, suggests that with native enzyme a conformational change accompanying aspartate association contributes significantly to rate limitation at t > 15 °C, but that catalytic steps become definitively slower below 15 °C. Model kinetics are derived to show that this change in mechanism at low temperature can force an ordered substrate binding system to produce exchange-rate patterns consistent with a random binding system with all exchange rates equal. The nonlinear Arrhenius plot also has important consequences for current theories of catalytic and regulatory mechanisms for this enzyme.  相似文献   

14.
Glycosylation affects the physical properties of proteins in a number of ways including solubility and aggregation behavior. To elucidate the mechanism underlying these effects, we have measured second virial coefficients (A2) of the heavily glycosylated pheniophora lycii phytase (Phy) and its enzymatically deglycosylated counterpart (dgPhy) in native and in denatured form by means of small angle x-ray scattering. The measured A2-values show that the native forms of Phy and dgPhy are equally repulsive at the studied pH 8 where A2 equals 10.9 ± 0.1 × 104 mL mol g−2. However, when thermally denatured, the A2 of dgPhy decreases to 9.0 ± 0.2 × 104 mL mol g−2 whereas it remained unchanged for Phy. In accord with earlier investigations, the p(r)-function measured here suggested that the glycans did not affect the peptide structure of the native protein. Conversely, glycosylation markedly changed the structure of thermally denatured protein. This was evident from the radius of gyration, which increased by 32% for Phy and only 11% for dgPhy on denaturation. We suggest that this expanding effect of the glycans on the denatured protein conformation relies on steric hindrance that limits the range of torsion angles available to the polypeptide.  相似文献   

15.
Protein amide hydrogen exchange (HDX) is a convoluted process, whose kinetics is determined by both dynamics of the protein and the intrinsic exchange rate of labile hydrogen atoms fully exposed to solvent. Both processes are influenced by a variety of intrinsic and extrinsic factors. A mathematical formalism initially developed to rationalize exchange kinetics of individual amide hydrogen atoms is now often used to interpret global exchange kinetics (e.g., as measured in HDX MS experiments). One particularly important advantage of HDX MS is direct visualization of various protein states by observing distinct protein ion populations with different levels of isotope labeling under conditions favoring correlated exchange (the so-called EX1 exchange mechanism). However, mildly denaturing conditions often lead to a situation where the overall HDX kinetics cannot be clearly classified as either EX1 or EX2. The goal of this work is to develop a framework for a generalized exchange model that takes into account multiple processes leading to amide hydrogen exchange, and does not require that the exchange proceed strictly via EX1 or EX2 kinetics. To achieve this goal, we use a probabilistic approach that assigns a transition probability and a residual protection to each equilibrium state of the protein. When applied to a small protein chymotrypsin inhibitor 2, the algorithm allows complex HDX patterns observed experimentally to be modeled with remarkably good fidelity. On the basis of the model we are now in a position to begin to extract quantitative dynamic information from convoluted exchange kinetics.  相似文献   

16.
Two main types of conformational fluctuations, local and global, are characteristic of the native protein structure and are detectable by hydrogen exchange. The probability of such fluctuations changes to a different degree during hemoglobin (Hb) oxygenation, changes in pH, and splitting of the intersubunit contracts. For comparison with the effect of heme removal, the rate of the hydrogen-deuterium (H-D) exchange of peptide H atoms (PHs) of human apoHb was studied by IR spectroscopy at pH 5.5–9.0 and temperatures of 10–38°C. The removal of heme increased the H-D exchange rate for 80% of Hb PHs with the exchange retardation factor P ~ 102-108. For the majority of PHs, the probability of local fluctuations depended weakly on the temperature; changes in enthalpy upon such local conformational transitions were ΔH op o = 0–15 kcal/M. Global fluctuations, displaying a stronger temperature dependence, did not arise with an increase in temperature to 38°C at pH 7.0, although apoHb began slowly denaturing and aggregating under these conditions. Destabilization of the apoHb structure with a concurrent decrease in pH to 5.5 and temperature to 10°C intensified global fluctuations in the native protein structure with ΔH op o < 0. The mechanism underlying the overall intensification of local fluctuations upon the heme removal, the specific features of apoHb heat denaturation under conditions close to those of in vivo Hb self-assembly, and the analogies between low-temperature global fluctuations and cold denaturation of globular proteins are discussed.  相似文献   

17.
In vivo room temperature chlorophyll a fluorescence coupled with CO2 and O2 exchange was measured to determine photosynthetic limitation(s) for spring and winter wheat (Triticum aestivum L.) grown at cold-hardening temperatures (5°C/5°C, day/night). Plants of comparable physiological stage, but grown at nonhardening temperatures (20°C/16°C, day/night) were used in comparison. Winter wheat cultivars grown at 5°C had light-saturated rates of CO2 exchange and apparent photon yields for CO2 exchange and O2 evolution that were equal to or greater than those of winter cultivars grown at 20°C. In contrast, spring wheat cultivars grown at 5°C showed 35% lower apparent photon yields for CO2 exchange and 25% lower light-saturated rates of CO2 exchange compared to 20°C grown controls. The lower CO2 exchange capacity is not associated with a lower efficiency of photosystem II activity measured as either the apparent photon yield for O2 evolution, the ratio of variable to maximal fluorescence, or the level of reduced primary quinone electron acceptor maintained at steady-state photosynthesis, and is most likely associated with carbon metabolism. The lower CO2 exchange capacity of the spring cultivars developed following long-term exposure to low temperature and did not occur following over-night exposure of nonhardened plants to 5°C.  相似文献   

18.
The effect of strongly destabilizing mutations, I106A and V108G of Ribonuclease A (RNase A), on its structure and stability has been determined by NMR. The solution structures of these variants are essentially equivalent to RNase A. The exchange rates of the most protected amide protons in RNase A (35°C), the I106A variant (35°C), and the V108G variant (10°C) yield stability values of 9.9, 6.0, and 6.8 kcal/mol, respectively, when analyzed assuming an EX2 exchange mechanism. Thus, the destabilization induced by these mutations is propagated throughout the protein. Simulation of RNase A hydrogen exchange indicates that the most protected protons in RNase A and the V108G variant exchange via the EX2 regime, whereas those of I106A exchange through a mixed EX1 + EX2 process. It is striking that a single point mutation can alter the overall exchange mechanism. Thus, destabilizing mutations joins high temperatures, high pH and the presence of denaturating agents as a factor that induces EX1 exchange in proteins. The calculations also indicate a shift from the EX2 to the EX1 mechanism for less protected groups within the same protein. This should be borne in mind when interpreting exchange data as a measure of local stability in less protected regions.  相似文献   

19.

Background and Aims

Deciduous trees have a seasonal carbon dioxide exchange pattern that is attributed to changes in leaf biochemical properties. However, it is not known if the pattern in leaf biochemical properties – maximum Rubisco carboxylation (Vcmax) and electron transport (Jmax) – differ between species. This study explored whether a general pattern of changes in Vcmax, Jmax, and a standardized soil moisture response accounted for carbon dioxide exchange of deciduous trees throughout the growing season.

Methods

The model MAESTRA was used to examine Vcmax and Jmax of leaves of five deciduous trees, Acer rubrum ‘Summer Red’, Betula nigra, Quercus nuttallii, Quercus phellos and Paulownia elongata, and their response to soil moisture. MAESTRA was parameterized using data from in situ measurements on organs. Linking the changes in biochemical properties of leaves to the whole tree, MAESTRA integrated the general pattern in Vcmax and Jmax from gas exchange parameters of leaves with a standardized soil moisture response to describe carbon dioxide exchange throughout the growing season. The model estimates were tested against measurements made on the five species under both irrigated and water-stressed conditions.

Key Results

Measurements and modelling demonstrate that the seasonal pattern of biochemical activity in leaves and soil moisture response can be parameterized with straightforward general relationships. Over the course of the season, differences in carbon exchange between measured and modelled values were within 6–12 % under well-watered conditions and 2–25 % under water stress conditions. Hence, a generalized seasonal pattern in the leaf-level physiological change of Vcmax and Jmax, and a standardized response to soil moisture was sufficient to parameterize carbon dioxide exchange for large-scale evaluations.

Conclusions

Simplification in parameterization of the seasonal pattern of leaf biochemical activity and soil moisture response of deciduous forest species is demonstrated. This allows reliable modelling of carbon exchange for deciduous trees, thus circumventing the need for extensive gas exchange experiments on different species.Key words: Carbon budget, deciduous trees, modelling, MAESTRA, soil moisture, species response, transpiration, Acer rubrum, Betula nigra, Quercus nuttallii, Q. phellos, Paulownia elongata  相似文献   

20.
《Bone and mineral》1994,24(3):223-234
To evaluate the effect of 17β-estradiol replacement (10 μg, twice a week) (E2) and treadmill exercise (18 m/min, 45 min/day) (EX) on long bone and vertebral bone mass and density, 10-month-old rats were ovariectomized (OV) and divided into four groups: OV, OV + E2, OV + EX, OV + EX + E2 2 months after surgery. After 7 weeks intervention, the calcium content and the density of lumbar-5 were higher in both OV + E2 and OV + EX + E2 groups than in the OV group, but, only the OV + EX + E2 group had a significantly higher femoral bone weight and density than the OV group. After 16 weeks intervention, the bone-conserving effects of E2 and EX were significant on lumbar-5 and femoral dry weight and density. The effect of E2 on both two sides of bones was due to the suppression of the bone turnover rate, while EX suppressed bone turnover rate primarily on the femur. We conclude that the effect of the two interventions on lumbar-5 and femoral bone mass were additive and independent.  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