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 609 毫秒
1.
A novel chiral sensor based on the self‐assembled monolayer of (6A‐ω‐mercaptoethylureado‐6A‐deoxy)heptakis(2,3‐di‐o‐phenylcarbamoyl)‐6B, 6C, 6D, 6E, 6F, 6G‐ hexa‐o‐phenylcarbamoyl‐β‐cyclodextrin (Ph‐β‐CD‐SH) on a quartz crystal transducer for chiral recognition was set up. (R,S)‐(±)‐(3‐Methoxyphenyl)ethylamine were recognized by this QCM chiral sensor with a QCM chiral discrimination factor of 1.33. Furthermore, UV spectroscopy was used to investigate the mechanism of host‐guest interactions between (6A‐azido‐6A‐deoxy)heptakis(2,3‐di‐o‐phenylcarbamoyl)‐6B, 6C, 6D, 6E, 6F, 6G‐hexa‐o‐phenylcarbamoyl‐β‐cyclodextrin (Ph‐β‐CD) and (R,S)‐(±)‐(3‐methoxyphenyl) ethylamine. The UV discrimination factor was determined to be 0.066. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

2.
The chiral discrimination studies of biological system are theoretically and practically significant for the development of chiral drugs and life science. Our work has embarked upon the interaction between serum albumin (SA) (including human SA and bovine SA), R,S‐1‐(4‐methoxyphenyl)ethylamine, and R,S‐1‐(3‐methoxyphenyl)ethylamine. The formation of intermediate transition state, binding sites, and chiral discrimination ability can be investigated by ultraviolet‐visible spectra and fluorescence spectra. Moreover, both the changes of hydrophobic microenvironment and energy transfer can be detected by synchronous fluorescence spectra and fluorescence lifetime. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

3.
Axially chiral biphenyls such as (M,S)‐ 3k have been conveniently obtained by crystallization of their diastereomeric mixtures, which were synthesized from racemic 4,4′‐dimethoxy‐5,6,5′,6′‐bis(methylenedioxy)‐2‐carboxylester‐2′‐carboxyl‐biphenyls 4 and chiral amino alcohols (R)‐alaninol, (S)‐alaninol, (S)‐valinol, and (S)‐phenylalaninol. A crystallization‐induced configuration transformation of the biphenyls was thus achieved. It was found that amide formation of an (S)‐valinol or (S)‐phenylalaninol at the 2′‐position of the biphenyl usually induced the deposition of crystals with the (M)‐configuration from ethanol in yields higher than 50%. The absolute configurations (ACs) of two crystalline biphenyls have been determined by X‐ray crystallographic analysis. The ACs of nine biphenyls have been assigned based on their CD spectra. Further, stability investigation of these axially chiral biphenyls revealed that the ACs could revert upon redissolution. The energy barrier to epimerization between (P,R)‐ 3b and (M,R)‐ 3b was measured as ΔG# = 21.45 kcal/mol and the half‐life in ethanol at 301 K was 17.1 h. Chirality, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

4.
Chiral considerations are found to be very much relevant in various aspects of forensic toxicology and pharmacology. In forensics, it has become increasingly important to identify the chirality of doping agents to avoid legal arguments and challenges to the analytical findings. The scope of this study was to develop an liquid chromatography–mass spectrometry (LCMS) method for the enantiomeric separation of typical illicit drugs such as ephedrines (ie, 1S,2R(+)‐ephedrine and 1R,2S(?)‐ephedrine) and pseudoephedrine (ie, R,R(?)‐pseudoephedrine and S,S(+)‐pseudoephedrine) by using normal phase chiral liquid chromatography–high‐resolution mass spectrometry technique. Results show that the Lux i‐amylose‐1 stationary phase has very broad and balancing‐enantio‐recognition properties towards ephedrine analogues, and this immobilized chiral stationary phase may offer a powerful tool for enantio‐separation of different types of pharmaceuticals in the normal phase mode. The type of mobile phase and organic modifier used appear to have dramatic influences on separation quality. Since the developed method was able to detect and separate the enantiomers at very low levels (in pico grams), this method opens easy access for the unambiguous identification of these illicit drugs and can be used for the routine screening of the biological samples in the antidoping laboratories.  相似文献   

5.
A novel nickel(II) hexaaza macrocyclic complex, [Ni(LR,R)](ClO4)2 ( 1 ), containing chiral pendant groups was synthesized by an efficient one‐pot template condensation and characterized (LR,R═1,8‐di((R)‐α‐methylnaphthyl)‐1,3,6,8,10,13‐hexaazacyclotetradecane). The crystal structure of compound 1 was determined by single‐crystal X‐ray analysis. The complex was found to have a square‐planar coordination environment for the nickel(II) ion. Open framework [Ni(LR,R)]3[C6H3(COO)3]2 ( 2 ) was constructed from the self‐assembly of compound 1 with deprotonated 1,3,5‐benzenetricarboxylic acid, BTC3?. Chiral discrimination of rac‐1,1′‐bi‐2‐naphthol and rac‐2,2,2‐trifluoro‐1‐(9‐anthryl)ethanol was performed to determine the chiral recognition ability of the chiral complex ( 1 ) and its self‐assembled framework ( 2 ). Binaphthol showed a good chiral discrimination on the framework ( 2 ). The optimum experimental conditions for the chiral discrimination were examined by changing the weight ratio between the macrocyclic complex 1 or self‐assembled framework 2 and racemates. The detailed synthetic procedures, spectroscopic data including single‐crystal X‐ray analysis, and the results of the chiral recognition for the compounds are described. Chirality, 25:54‐58, 2013 © 2012 Wiley Periodicals, Inc.  相似文献   

6.
We are reporting the synthesis, characterization, and calf thymus DNA binding studies of novel chiral macrocyclic Mn(III) salen complexes S‐1 , R‐1 , S‐2 , and R‐2 . These chiral complexes showed ability to bind with DNA, where complex S‐1 exhibits the highest DNA binding constant 1.20 × 106 M?1. All the compounds were screened for superoxide and hydroxyl radical scavenging activities; among them, complex S‐1 exhibited significant activity with IC50 1.36 and 2.37 μM, respectively. Further, comet assay was used to evaluate the DNA damage protection in white blood cells against the reactive oxygen species wherein complex S‐1 was found effective in protecting the hydroxyl radicals mediated plasmid and white blood cells DNA damage. Chirality 24:1063–1073, 2012.© 2012 Wiley Periodicals, Inc.  相似文献   

7.
The interaction of the nonsteroidal anti‐inflammatory drug flurbiprofen (FBP) with human serum albumin (HSA) hardly influences the fluorescence of the protein's single tryptophan (Trp). Therefore, in addition to fluorescence, heavy atom‐induced room‐temperature phosphorescence is used to study the stereoselective binding of FBP enantiomers and their methyl esters to HSA. Maximal HSA phosphorescence intensities were obtained at a KI concentration of 0.2 M. The quenching of the Trp phosphorescence by FBP is mainly dynamic and based on Dexter energy transfer. The Stern–Volmer plots based on the phosphorescence lifetimes indicate that (R)‐FBP causes a stronger Trp quenching than (S)‐FBP. For the methyl esters of FBP, the opposite is observed: (S)‐(FBPMe) quenches more than (R)‐FBPMe. The Stern–Volmer plots of (R)‐FBP and (R)‐FBPMe are similar although their high‐affinity binding sites are different. The methylation of (S)‐FBP causes a large change in its effect on the HSA phosphorescence lifetime. Furthermore, the quenching constants of 3.0 × 107 M?1 s?1 of the R‐enantiomers and 2.5 × 107 M?1 s?1 for the S‐enantiomers are not influenced by the methylation and indicate a stereoselectivity in the accessibility of the HSA Trp to these drugs. Chirality 24:840–846, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

8.
The presystemic sulfate conjugation of the stereoisomers of 4′‐methoxyfenoterol, (R,R′)‐MF, (S,S′)‐MF, (R,S′)‐MF, and (S,R′)‐MF, was investigated using commercially available human intestinal S9 fractions, a mixture of sulfotransferase (SULT) enzymes. The results indicate that the sulfation was stereospecific and that an S‐configuration at the β‐OH carbon of the MF molecule enhanced the maximal formation rates with (S,R′)‐MF (S,S′)‐MF (R,S′)‐MF ≈ (R,R′)‐MF, and competition studies demonstrated that (S,R′)‐MF is an effective inhibitor of (R,R′)‐MF sulfation (IC50 = 60 μM). In addition, the results from a cDNA‐expressed human SULT isoform screen indicated that SULT1A1, SULT1A3, and SULT1E1 can mediate the sulfation of all four MF stereoisomers. Previously published molecular models of SULT1A3 and SULT1A1 were used in docking simulations of the MF stereoisomers using Molegro Virtual Docker. The models of the MF‐SULT1A3 and MF‐SULT1A1 complexes indicate that each of the two chiral centers of MF molecule plays a role in the observed relative stabilities. The observed stereoselectivity is the result of multiple hydrogen bonding interactions and induced conformational changes within the substrate–enzyme complex. In conclusion, the results suggest that a formulation developed from a mixture of (R,R′)‐MF and (S,R′)‐MF may increase the oral bioavailability of (R,R′)‐MF. Chirality 24:796–803, 2012. © 2012 Wiley Periodicals, Inc. 1   相似文献   

9.
Luminescent lanthanide (III) ions have been exploited for circularly polarized luminescence (CPL) for decades. However, very few of these studies have involved chiral samarium (III) complexes. Complexes are prepared by mixing axial chiral ligands (R/S))‐2,2’‐bis(diphenylphosphoryl)‐1,1′‐binaphthyl (BINAPO) with europium and samarium Tris (trifluoromethane sulfonate) (Eu (OTf)3 and Sm (OTf)3). Luminescence‐based titration shows that the complex formed is Ln((R/S)‐BINAPO)2(OTf)3, where Ln = Eu or Sm. The CPL spectra are reported for Eu((R/S)‐BINAPO)2(OTf)3 and Sm((R/S)‐BINAPO)2(OTf)3. The sign of the dissymmetry factors, gem, was dependent upon the chirality of the BINAPO ligand, and the magnitudes were relatively large. Of all of the complexes in this study, Sm((S)‐BINAPO)2(OTf)3 has the largest gem = 0.272, which is one of the largest recorded for a chiral Sm3+ complex. A theoretical three‐dimensional structural model of the complex that is consistent with the experimental observations is developed and refined. This report also shows that (R/S)‐BINAPO are the only reported ligands where gem (Sm3+) > gem (Eu3+).  相似文献   

10.
Phthalides and their precursors have demonstrated a large variety of biological activities. Eighteen phthalides were synthesized and tested on the stored grain pest Rhyzopertha dominica. In the screening bioassay, compounds rac‐(2R,2aS,4R,4aS,6aR,6bS,7R)‐7‐bromohexahydro‐2,4‐methano‐1,6‐dioxacyclopenta[cd]pentalen‐5(2H)‐one ( 15 ) and rac‐(3R,3aR,4R,7S,7aS)‐3‐(propan‐2‐yloxy)hexahydro‐4,7‐methano‐2‐benzofuran‐1(3H)‐one ( 17 ) showed mortality similar to the commercial insecticide, Bifenthrin® (≥90 %). The time (LT50) and dose (LD50) necessary to kill 50 % of the R. dominica population were determined for the most efficacious phthalides 15 and 17 . Compound 15 presented the lowest LD50 (1.97 μg g?1), being four times more toxic than Bifenthrin® (LD50=9.11 μg g?1). Both compounds presented an LT50 value equal to 24 h. When applied at a sublethal dose, both phthalides (especially compound 15 ), reduced the emergence of the first progeny of R. dominica. These findings highlight the potential of phthalides 15 and 17 as precursors for the development of insecticides for R. dominica control.  相似文献   

11.
Accessible chiral syntheses of 3 types of (R)‐2‐sulfanylcarboxylic esters and acids were performed: (R)‐2‐sulfanylpropanoic (thiolactic) ester (53%, 98%ee) and acid (39%, 96%ee), (R)‐2‐sulfanylsucciinic diester (59%, 96%ee), and (R)‐2‐mandelic ester (78%, 90%ee) and acid (59%, 96%ee). The present practical and robust method involves (i) clean SN2 displacement of methanesulfonates of (S)‐2‐hydroxyesters by using commercially available AcSK with tris(2‐[2‐methoxyethoxy])ethylamine and (ii) sufficiently mild deacetylation. The optical purity was determined by the corresponding (2R,5R)‐trans‐thiazolidin‐4‐one and (2S,5R)‐cis‐thiazolidin‐4‐one derivatives based on accurate high‐performance liquid chromatography analysis with high‐resolution efficiency. Compared with the reported method utilizing AcSCs (generated from AcSH and CsCO3), the present method has several advantages, that is, the use of odorless AcCOSK reagent, reasonable reaction velocity, isolation procedure, and accurate, reliable optical purity determination. The use of accessible AcSK has advantages because of easy‐to‐handle odorless and hygroscopic solid that can be used in a bench‐top procedure. The Ti(OiPr)4 catalyst promoted smooth trans‐cyclo‐condensation to afford (2R,5R)‐trans‐thiazolidin‐4‐one formation of (R)‐2‐sulfanylcarboxylic esters with available N‐(benzylidene)methylamine under neutral conditions without any racemization, whereas (2S,5R)‐cis‐thiazollidin‐4‐ones were obtained via cis‐cyclo‐condensation and no catalysts. Direct high‐performance liquid chromatography analysis of methyl (R)‐mandelate was also performed; however, the resolution efficiency was inferior to that of the thaizolidin‐4‐one derivatizations.  相似文献   

12.
We analyzed 17 months (August 2005 to December 2006) of continuous measurements of soil CO2 efflux or soil respiration (RS) in an 18‐year‐old west‐coast temperate Douglas‐fir stand that experienced somewhat greater than normal summertime water deficit. For soil water content at the 4 cm depth (θ) > 0.11 m3 m?3 (corresponding to a soil water matric potential of ?2 MPa), RS was positively correlated to soil temperature at the 2 cm depth (TS). Below this value of θ, however, RS was largely decoupled from TS, and evapotranspiration, ecosystem respiration and gross primary productivity (GPP) began to decrease, dropping to about half of their maximum values when θ reached 0.07 m3 m?3. Soil water deficit substantially reduced RS sensitivity to temperature resulting in a Q10 significantly < 2. The absolute temperature sensitivity of RS (i.e. dRS/dTS) increased with θ up to 0.15 m3 m?3, above which it slowly declined. The value of dRS/dTS was nearly 0 for θ < 0.08 m3 m?3, thereby confirming that RS was largely unaffected by temperature under soil water stress conditions. Despite the possible effects of seasonality of photosynthesis, root activity and litterfall on RS, the observed decrease in its temperature sensitivity at low θ was consistent with the reduction in substrate availability due to a decrease in (a) microbial mobility, and diffusion of substrates and extracellular enzymes, and (b) the fraction of substrate that can react at high TS, which is associated with low θ. We found that an exponential (van't Hoff type) model with Q10 and R10 dependent on only θ explained 92% of the variance in half‐hourly values of RS, including the period with soil water stress conditions. We hypothesize that relating Q10 and R10 to θ not only accounted for the effects of TS on RS and its temperature sensitivity but also accounted for the seasonality of biotic (photosynthesis, root activity, and litterfall) and abiotic (soil moisture and temperature) controls and their interactions.  相似文献   

13.
Two new sterols 1 and 2 and five known ones 3 – 7 were isolated for the first time from the fruiting bodies of Cortinarius glaucopus. Their structures were established by 1‐ and 2D‐NMR spectra and HR‐FABS‐MS. The relative configuration of 1 was firmly determined by comparison of the observed 1H–1H couplings and NOESY correlations, with those predicted for the computed geometries of the conformers. Calculations were performed by means of DFT with the B3LYP functional at 6‐31 + G(d,p) level of theory, in CHCl3 as the solvent. The structures of the new ergosterol derivatives, called glaucoposterol A ( 1 ) and B ( 2 ), were thus established as (3S,5R,7R,10R,13R,17R,20S,22R,23R,24R)‐5,6‐epoxy‐3,7,23‐trihydroxystrophast‐8‐en‐14‐one and (22E,3S,5S,9S,10R,13R,17R,20R,24R)‐3,5‐dihydroxyergosta‐6,8(14),22‐trien‐15‐one, respectively. Moreover, the configuration of known strophasterol C ( 3 ) was determined as (3S,5R,6S,7R,10R,13R,17R,20S,22S,24R). Glaucoposterol A ( 1 ) and strophasterol C ( 3 ) represent the second finding in nature of steroids with the rare strophastane skeleton.  相似文献   

14.
The pyrrolidine side chain makes proline play a unique role in protein structure and function. The Cγ ring pucker preference and the cis trans peptidyl bond ratio can be mediated via stereoelectronic effects. Here we used a compact triple‐stranded antiparallel β‐sheet protein, the human Pin1 WW domain, to study the consequences of implanting a preorganized Cγ ring pucker on protein structure and function. The conserved Pro37 is a key residue involved in one hydrophobic core, plays an important role in the WW domain, and adopts a Cγendo ring pucker in the native structure. Pro37 was replaced with Cγexo biased pucker derivatives: (2S,4R)‐4‐hydroxyproline (4R‐Hyp), (2S,4R)‐4‐fluoroproline (4R‐Flp), (2S,4R)‐4‐methoxyproline (4R‐Mop), and Cγendo biased pucker derivatives: (2S,4S)‐4‐hydroxyproline (4S‐hyp), (2S,4S)‐4‐fluoroproline (4S‐flp), (2S,4S)‐4‐methoxyproline (4S‐mop) to examine how a preorganized pucker affects the folding stability and ligand‐binding affinity. Circular dichroism measurements indicate that among the variants, only the one with 4S‐flp substitution (P37flp) is more stable than the wild type, suggesting that the stabilization effects originated from preorganization of the backbone conformation and the hydrophobicity of C? F group. Analysis of ligand‐binding affinity using isothermal titration calorimetry revealed that only P37flp has a stronger ligand affinity than the wild type, showing that 4S‐flp can stabilize the WW domain and increase its ligand affinity. Together we have used 4‐substituted proline derivatives and the WW domain to demonstrate that proline ring puckering can be a key factor in determining the folding stability of a protein but the choice of the derivative groups is also critical. Proteins 2014; 82:67–76. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
This study investigated the impact of predicted future climatic and atmospheric conditions on soil respiration (RS) in a Danish Calluna‐Deschampsia‐heathland. A fully factorial in situ experiment with treatments of elevated atmospheric CO2 (+130 ppm), raised soil temperature (+0.4 °C) and extended summer drought (5–8% precipitation exclusion) was established in 2005. The average RS, observed in the control over 3 years of measurements (1.7 μmol CO2 m?2 sec?1), increased 38% under elevated CO2, irrespective of combination with the drought or temperature treatments. In contrast, extended summer drought decreased RS by 14%, while elevated soil temperature did not affect RS overall. A significant interaction between elevated temperature and drought resulted in further reduction of RS when these treatments were combined. A detailed analysis of short‐term RS dynamics associated with drought periods showed that RS was reduced by ~50% and was strongly correlated with soil moisture during these events. Recovery of RS to pre‐drought levels occurred within 2 weeks of rewetting; however, unexpected drought effects were observed several months after summer drought treatment in 2 of the 3 years, possibly due to reduced plant growth or changes in soil water holding capacity. An empirical model that predicts RS from soil temperature, soil moisture and plant biomass was developed and accounted for 55% of the observed variability in RS. The model predicted annual sums of RS in 2006 and 2007, in the control, were 672 and 719 g C m?2 y?1, respectively. For the full treatment combination, i.e. the future climate scenario, the model predicted that soil respiratory C losses would increase by ~21% (140–150 g C m?2 y?1). Therefore, in the future climate, stimulation of C storage in plant biomass and litter must be in excess of 21% for this ecosystem to not suffer a reduction in net ecosystem exchange.  相似文献   

16.
This study provides the first measurements of the standard respiration rate (RS) and growth dynamics of European sardine Sardina pilchardus larvae reared in the laboratory. At 15° C, the relationship between RS (µl O2 individual?1 h?1) and larval dry mass (MD, µg) was equal to: RS = 0·0057(±0·0007, ± s.e.)·MD0·8835(±0·0268), (8–11% MD day?1). Interindividual differences in RS were not related to interindividual differences in growth rate or somatic (Fulton's condition factor) or biochemical‐based condition (RNA:DNA).  相似文献   

17.
Paclobutrazol, with two stereogenic centers, but gives only (2R, 3R) and (2S, 3S)‐enantiomers because of steric‐hindrance effects, is an important plant growth regulator in agriculture and horticulture. Enantioselective degradation of paclobutrazol was investigated in rat liver microsomes in vitro. The degradation kinetics and the enantiomer fraction were determined using a Lux Cellulose‐1 chiral column on a reverse‐phase liquid chromatography–tandem mass spectrometry system. The t1/2 of (2R, 3R)‐paclobutrazol is 18.60 min, while the t1/2 of (2S, 3S)‐paclobutrazol is 10.93 min. Such consequences clearly indicated that the degradation of paclobutrazol in rat liver microsomes was stereoselective and the degradation rate of (2S, 3S)‐paclobutrazol was much faster than (2R, 3R)‐paclobutrazol. In addition, significant differences between the two enantiomers were also observed in enzyme kinetic parameters. The Vmax of (2S, 3S)‐paclobutrazol was more than 2‐fold of (2R, 3R)‐paclobutrazol and the Clint of (2S, 3S)‐paclobutrazol was higher than that of (2R, 3R)‐paclobutrazol after incubation in rat liver microsomes. These results may have potential implications for better environmental and ecological risk assessment for paclobutrazol. Chirality 27:344–348, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

18.
Compounds based on the pyrroloquinoxaline system can interact with serotonin 5‐HT3, cannabinoid CB1, and μ‐opioid receptors. Herein, a chiral pool synthesis of diastereomerically and enantiomerically pure bromolactam (S,R,R,R)‐ 14A is presented. Introduction of the cyclohexenyl ring at the N‐atom of (S)‐proline derivatives 8 or methyl (S)‐pyroglutamate ( 12 ) led to the N‐cyclohexenyl substituted pyrrolidine derivatives 4 and 13 , respectively. All attempts to cyclize the (S)‐proline derivatives 4 with a basic pyrrolidine N‐atom via [3 + 2] cycloaddition, aziridination, or bromolactamization failed. Fast aromatization occurred during treatment of cyclohexenamines under halolactamization conditions. In contrast, reaction of a 1:1 mixture of diastereomeric pyroglutamates (S,R)‐ 13bA and (S,S)‐ 13bB with LiOtBu and NBS provided the tricyclic bromolactam (S,R,R,R)‐ 14A with high diastereoselectivity from (S,R)‐ 13bA , but did not transform the diastereomer (S,S)‐ 13bB . The different behavior of the diastereomeric pyroglutamates (S,R)‐ 13bA and (S,S)‐ 13bB is explained by different energetically favored conformations. Chirality 26:793–800, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

19.
In‐depth conformational analyses of 10 known eremophilane (= (1S,4aR,7R,8aR)‐decahydro‐1,8a‐dimethyl‐7‐(1‐methylethyl)napththalene) sesquiterpenes, 1 – 10 , from Petasites hybridus were performed with molecular mechanics as well as density functional theory methods. Electronic transition energies and rotational strengths of these eight eremophilane lactones and two petasins were calculated by time‐dependent density functional theory (B3PW91/TZVP). The absolute configurations of the constituents could be assigned by comparison of their simulated and experimental circular dichroism (CD) spectra in methanol as (4S,5R,8S,10R) ( 1 , 2 ), (2R,4S,5R,8S,10R) ( 3 , 4 , 5 ), (2R,4S,5R,8R,9R,10R) ( 6 ), (2R,4S,5R,8R,10R) ( 7 , 8 ), and (3R,4R,5R) ( 9 , 10 ). Single‐crystal X‐ray diffraction data of 8β‐hydroxyeremophilanolide ((8S)‐8‐hydroxyeremophil‐7(11)‐en‐12,8‐olide) ( 1 ) served as starting point for the theoretical conformational calculations of the 8β‐epimers of the eremophilane lactones. Experimental CD spectra as well as 1H NMR spectra of compound 1 in methanol were considerably dependent on sample concentration. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

20.
We used estimates of autotrophic respiration (RA), net primary productivity (NPP) and soil CO2 evolution (Sff), to develop component carbon budgets for 12‐year‐old loblolly pine plantations during the fifth year of a fertilization and irrigation experiment. Annual carbon use in RA was 7.5, 9.0, 15.0, and 15.1 Mg C ha?1 in control (C), irrigated (I), fertilized (F) and irrigated and fertilized (IF) treatments, respectively. Foliage, fine root and perennial woody tissue (stem, branch, coarse and taproot) respiration accounted for, respectively, 37%, 24%, and 39% of RA in C and I treatments and 38%, 12% and 50% of RA in F and IF treatments. Annual gross primary production (GPP=NPP+RA) ranged from 13.1 to 26.6 Mg C ha?1. The I, F, and IF treatments resulted in a 21, 94, and 103% increase in GPP, respectively, compared to the C treatment. Despite large treatment differences in NPP, RA, and carbon allocation, carbon use efficiency (CUE=NPP/GPP) averaged 0.42 and was unaffected by manipulating site resources. Ecosystem respiration (RE), the sum of Sff, and above ground RA, ranged from 12.8 to 20.2 Mg C ha?1 yr?1. Sff contributed the largest proportion of RE, but the relative importance of Sff decreased from 0.63 in C treatments to 0.47 in IF treatments because of increased aboveground RA. Aboveground woody tissue RA was 15% of RE in C and I treatments compared to 25% of RE in F and IF treatments. Net ecosystem productivity (NEP=GPP‐RE) was roughly 0 in the C and I treatments and 6.4 Mg C ha?1 yr?1 in F and IF treatments, indicating that non‐fertilized treatments were neither a source nor a sink for atmospheric carbon while fertilized treatments were carbon sinks. In these young stands, NEP is tightly linked to NPP; increased ecosystem carbon storage results mainly from an increase in foliage and perennial woody biomass.  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