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 31 毫秒
1.
Rapid penetration of baby hamster kidney cells was used as a selective pressure for the isolation of pathogenesis mutants of the S.A.AR86 strain of Sindbis virus. Unlike most Sindbis virus strains, S.A.AR86 is virulent in adult as well as neonatal mice. Two classes of mutants were defined. One class was attenuated in adult mice inoculated intracerebrally as well as in neonatal mice inoculated either intracerebrally or subcutaneously. Sequence analysis of the glycoprotein genes of the parent virus and three such mutant strains revealed a single point mutation which resulted in an amino acid change at position 1 in the E2 glycoprotein. The change from a serine in S.A.AR86 to an asparagine in the mutants created a new site for N-linked glycosylation which appeared to be utilized. This mutation did not retard release of infectious particles; however, mutant virions contained the E2 precursor protein (PE2) rather than the E2 glycoprotein itself. The mutants also lost the ability to bind two E2-specific monoclonal antibodies, R6 and R13. A second class of mutants was attenuated in neonatal mice upon subcutaneous inoculation but remained virulent in adults and in neonates when inoculated intracerebrally. Sequence analysis of three such strains revealed the substitution of an arginine residue for a serine at position 114 in the E2 glycoprotein. Reactivity with monoclonal antibodies R6 and R13 was reduced, yet members of this mutant class were more susceptible than S.A.AR86 to neutralization by these antibodies.  相似文献   

2.
Alanine substitutions were made for 15 amino acids in the cytoplasmic loop between transmembrane helices 6 and 7 (L6/7) of the human alpha(1)-subunit of Na,K-ATPase. Most mutations reduced Na,K-ATPase activity by less than 50%; however, the mutations R834A, R837A, and R848A reduced Na,K-ATPase activity by 75, 89, and 66%, respectively. Steady-state phosphoenzyme formation from ATP was reduced in mutants R834A, R837A, and R848A, and R837A also had a faster E(2)P --> E(2) dephosphorylation rate compared with the wild-type enzyme. Effects of L6/7 mutations on the phosphorylation domain of the protein were also demonstrated by (18)O exchange, which showed that intrinsic rate constants for P(i) binding and/or reaction with the protein were altered. Although most L6/7 mutations had no effect on the interaction of Na(+) or K(+) with Na,K-ATPase, the E825A, E828A, R834A, and R837A mutations reduced the apparent affinity of the enzyme for both Na(+) and K(+) by 1.5-3-fold. 1-Bromo-2,4,6-tris(methylisothiouronium)benzene (Br-TITU(3+)), a competitive antagonist of Rb(+) and Na(+) occlusion, was used to test whether charged residues in L6/7 are involved in binding monovalent cations and cation antagonists. Br-TITU(3+) inhibited ouabain binding to wild type Na,K-ATPase with an IC(50) of 30 microM. Ouabain binding to the E825A, E828A, R834A, or R837A mutants was still inhibited by Br-TITU(3+), indicating that Br-TITU(3+) does not bind to charged residues in L6/7. This observation makes it unlikely that L6/7 functions as a cytoplasmic cation binding site in Na,K-ATPase, and together with the effects of L6/7 mutations on phosphate interactions with the enzyme suggests that L6/7 is important in stabilizing the phosphorylation domain and its relationship to the ion binding sites of the protein.  相似文献   

3.
ABSTRACT Between May 1979 and August 1991, 48.7% (57/117) of the harvest mice ( Reithrodontomys spp.) examined from 10 localities in Mexico, California and New Mexico had coccidian oocysts in their feces. A total of 46.7% (49/105) of the Reithrodontomys megalotis examined were positive for coccidian oocysts; this included samples from five states in Mexico (47.1%, 8/17), three counties in California (66.7%, 4/6) and two counties in New Mexico (45.1%, 37/82); 66.7% (8/12) of the Reithrodontomys montanus from one county in New Mexico also were infected. Only two coccidian species, Eimeria arizonensis and Eimeria langebarteli , were found in these hosts. Oocysts of E. langebarteli were found only in R. megalotis : in all three infected mice from Madera County, California, in the only mouse from San Bernardino County, California, and in 63% (5/8) of the infected mice from four states in Mexico. Oocysts of E. arizonensis were found in R. megalotis in Mexico, California, and New Mexico and in R. montanus from New Mexico. Sporulated oocysts of E. langebarteli differed slightly from those in previously published reports by having wider oocysts and larger sporocysts. Sporulated oocysts of E. arizonensis were variable in size, with those recovered from R. montanus significantly larger in length and width and sporocyst width than those from R. megalotis . The structure of the oocyst residuum was polymorphic, both within and between host species, and within the same mouse; it could appear as one large globule, two globules, several to many smaller globules, or as a compact mass of many small granules. Oocysts with a variable residuum were larger than those with one globule in all oocyst/sporocyst dimensions. Only 9% (5/57) of the infected mice were discharging oocysts of both eimerians when examined.  相似文献   

4.
5.
Genetic defects in pyruvate dehydrogenase complex (PDC) cause lactic acidosis, neurological deficits, and often early death. Most mutations of PDC are localized in the alpha subunit of the pyruvate dehydrogenase (E1) component. We have kinetically characterized a patient's missense mutation alphaH44R in E1alpha by creating and purifying three recombinant human E1s (alphaH44R, alphaH44Q, and alphaH44A). Substitutions at histidine-15 resulted in decreased V(max) values (6% alphaH44R; 30% alphaH44Q; 90% alphaH44A) while increasing K(m) values for thiamine pyrophosphate (TPP) compared to wild-type (alphaH44R, 3-fold; alphaH44Q, 7-fold; alphaH44A, 10-fold). This suggests that the volume of the residue at site 15 is important for TPP binding and substitution by a residue with a longer side chain disrupts the active site more than the TPP binding site. The rates of phosphorylation and dephosphorylation of alphaH44R E1 by E1-kinase and phospho-E1 phosphatase, respectively, were similar to that of the wild-type E1 protein. These results provide a biochemical basis for altered E1 function in the alphaH44R E1 patient.  相似文献   

6.
The side chains of Escherichia coli phosphofructokinase (EcPFK) that interact with bound substrate, fructose 6-phosphate (Fru-6-P), are examined for their potential roles in allosteric regulation. Mutations that severely decrease Fru-6-P affinity and/or k(cat)/K(m) were created at each contact residue, with the exception of the catalytic base, D127. Even though Fru-6-P affinity was greatly decreased for R162E, M169A, E222A/H223A, and R243E, the mutated proteins retained the ability to be activated by MgADP and inhibited by phosphoenolpyruvate (PEP). R252E did not show an allosteric response to either MgADP or PEP. The H249E mutation retained MgADP activation but did not respond to PEP. R72E, T125A, and R171E maintained allosteric inhibition by PEP. Both R72E and T125A displayed a MgADP-dependent decrease in k(cat) but no MgADP-dependent K-type effects. R171E maintained MgADP-dependent K-type activation but also displayed a MgADP-dependent decrease in k(cat). Localization of mutations that alter MgADP activation near the transferred phosphate group indicates the importance of the 1-methoxy region of Fru-6-P in allosteric regulation by MgADP. A region near the 6'-phosphate may be similarly important for PEP inhibition. R252 is uniquely positioned between the 1'- and 6'-phosphates of bound Fru-1,6-BP, and the mutation at this position may alter both allosterically responsive regions. The differential functions of specific regions in the Fru-6-P contact residues support different mechanisms for allosteric activation and inhibition. In addition, the lack of correlation between mutations that decrease Fru-6-P affinity and those that abolish allosteric communications supports the independence of affinity and allosteric coupling.  相似文献   

7.
Hui EK  Barman S  Yang TY  Nayak DP 《Journal of virology》2003,77(12):7078-7092
Influenza type A virus matrix (M1) protein possesses multiple functional motifs in the helix 6 (H6) domain (amino acids 91 to 105), including nuclear localization signal (NLS) (101-RKLKR-105) involved in translocating M1 from the cytoplasm into the nucleus. To determine the role of the NLS motif in the influenza virus life cycle, we mutated these and the neighboring sequences by site-directed mutagenesis, and influenza virus mutants were generated by reverse genetics. Our results show that infectious viruses were rescued by reverse genetics from all single alanine mutations of amino acids in the H6 domain and the neighboring region except in three positions (K104A and R105A within the NLS motif and E106A in loop 6 outside the NLS motif). Among the rescued mutant viruses, R101A and R105K exhibited reduced growth and small-plaque morphology, and all other mutant viruses showed the wild-type phenotype. On the other hand, three single mutations (K104A, K105A, and E106A) and three double mutations (R101A/K102A, K104A/K105A, and K102A/R105A) failed to generate infectious virus. Deletion (Delta YRKL) or mutation (4A) of YRKL also abolished generation of infectious virus. However, replacement of the YRKL motif with PTAP or YPDL as well as insertion of PTAP after 4A mutation yielded infectious viruses with the wild-type phenotype. Furthermore, mutant M1 proteins (R101A/K102A, Delta YRKL, 4A, PTAP, 4A+PTAP, and YPDL) when expressed alone from cloned cDNAs were only cytoplasmic, whereas the wild-type M1 expressed alone was both nuclear and cytoplasmic as expected. These results show that the nuclear translocation function provided by the positively charged residues within the NLS motif does not play a critical role in influenza virus replication. Furthermore, these sequences of H6 domain can be replaced by late (L) domain motifs and therefore may provide a function similar to that of the L domains of other negative-strand RNA and retroviruses.  相似文献   

8.
《Biophysical journal》2022,121(9):1704-1714
In epithelial tumors, oncoprotein E6 binds with the ubiquitin ligase E6AP to form E6/E6AP heterodimer; then this heterodimer recruits p53 to form E6/E6AP/p53 heterotrimer and induces p53 degradation. Recent experiments demonstrated that three E6 single-site mutants (F47R, R102A, and L50E) can inhibit the E6/E6AP/p53 heterotrimer formation and rescue p53 from the degradation pathway. However, the molecular mechanism underlying mutation-induced heterotrimer inhibition remains largely elusive. Herein, we performed extensive molecular dynamics simulations (totally ~13 μs) on both heterodimer and heterotrimer to elucidate at an atomic level how each p53-degradation-defective HPV16 E6 mutant reduces the structural stabilities of the two complexes. Our simulations reveal that the three E6 mutations destabilize the structure of E6/E6AP/p53 complex through distinct mechanisms. Although F47RE6 mutation has no effect on the structure of E6/E6AP heterodimer, it results in an electrostatic repulsion between R47E6 and R290p53, which is unfavorable for E6-p53 binding. R102AE6 mutation destabilizes the structure of E6/E6AP heterodimer and significantly disrupts hydrophobic and cation-π interactions between F47E6 and E286p53/L298p53/R290p53. L50EE6 mutation impairs both E6 interdomain interactions (especially F47-K108 cation-π interaction) and E6-E6AP intermolecular interactions important for the stabilization of E6/E6AP heterodimer. This study identifies the intra- and intermolecular interactions crucial for the complex stability, which may provide mechanistic insights into the inhibition of complex formation by the three HPV16 E6 mutations.  相似文献   

9.
A collection of 56 purified thrombin mutants, in which 76 charged or polar surface residues on thrombin were mutated to alanine, was used to identify key residues mediating the interactions of thrombin with thrombomodulin (TM), protein C, and thrombin-activatable fibrinolysis inhibitor (TAFI). Comparison of protein C activation in the presence and absence of TM identified 11 residues mediating the thrombin-TM interaction (Lys(21), Gln(24), Arg(62), Lys(65), His(66), Arg(68), Thr(69), Tyr(71), Arg(73), Lys(77), Lys(106)). Three mutants (E25A, D51A, R89A/R93A/E94A) were found to have decreased ability to activate TAFI yet retained normal protein C activation, whereas three other mutants (R178A/R180A/D183A, E229A, R233A) had decreased ability to activate protein C but maintained normal TAFI activation. One mutant (W50A) displayed decreased activation of both substrates. Mapping of these functional residues on thrombin revealed that the 11 residues mediating the thrombin-TM interaction are all located in exosite I. Residues important in TAFI activation are located above the active-site cleft, whereas residues involved in protein C are located below the active-site cleft. In contrast to the extensive overlap of residues mediating TM binding and fibrinogen clotting, these data show that distinct domains in thrombin mediate its interactions with TM, protein C, and TAFI. These studies demonstrate that selective enzymatic properties of thrombin can be dissociated by site-directed mutagenesis.  相似文献   

10.
Y Snitko  S K Han  B I Lee  W Cho 《Biochemistry》1999,38(24):7803-7810
To identify the residues essential for interfacial binding and substrate binding of human pancreatic phospholipase A2 (hpPLA2), several ionic residues in the putative interfacial binding surface (R6E, K7E, K10E, and K116E) and substrate binding site (D53K and K56E) were mutated. Interfacial affinity of these mutants was measured using anionic polymerized liposomes, and their enzymatic activity was measured using various substrates including phospholipid monomers, zwitterionic and anionic micelles, and anionic polymerized mixed liposomes. Similar mutations (R6E, K10E, K56E, and K116E) were made to porcine pancreatic phospholipase A2 (ppPLA2), and the properties of mutants were measured by the same methods. Results indicate that hpPLA2 and ppPLA2 have similar interfacial binding mechanisms in which cationic residues in the amino terminus and Lys-116 in the carboxy terminus are involved in binding to anionic lipid surfaces. Small but definite differences between the two enzymes were observed in overall interfacial affinity and activity and the effects of the mutations on interfacial enzyme activity. The interfacial binding of hpPLA2 and ppPLA2 is distinct from that of bovine pancreatic phospholipase A2 in that Lys-56 is involved in the interfacial binding of the latter enzyme. The unique phospholipid headgroup specificity of hpPLA2 derives from the presence of Asp-53 in the substrate binding site. This residue appears to participate in stabilizing electrostatic interactions with the cationic ethanolamine headgroup, hence the phosphatidylethanolamine preference of hpPLA2. Taken together, these studies reveal the similarities and the differences in the mechanisms by which mammalian pancreatic phospholipases A2 interact with lipid aggregates and perform interfacial catalysis.  相似文献   

11.
12.
Domain I of the Cry1Ab insecticidal toxic protein has seven alpha-helices and is considered to be involved in the ion channel activity. While other alpha-helices, particularly alpha-4 and alpha-5, have been extensively explored, the remaining alpha-helices have been slightly studied. Site-directed mutagenesis was used to generate mutations throughout sequences encoding the alpha-helix 7 to test its role in ion channel function. Every amino acid residue in alpha-helix 7 was mutated to alanine. Most resultant proteins, e.g., D225A, W226A, Y229A, N230A, R233A, R234A, D242A, and F247A yielded no protoxin or were sensitive to degradation by trypsin or Manduca sexta midgut juice. Other mutant proteins, R224A, R228A, and E235A, were resistant to degradation to the above proteases but were 8, 30, and 12 times less toxic to M. sexta, respectively, than the wild-type Cry1Ab. Circular dichroism spectroscopy indicated a very small change in the R228A spectrum, while R224A and E235A display the same spectrum as the wild-type protein. These three mutant proteins showed little differences from Cry1Ab when analyzed by saturation binding and competition binding kinetics with (125)I-labeled toxin or by surface plasmon resonance to M. sexta brush border membrane vesicles. More conservative amino acid substitutions were introduced into alpha-helix 7 residues: R228K, F232Y, E235Q, and F247Y. In comparison with wild-type Cry1Ab, mutant proteins R228K, F232Y, E235A, and E235Q selectively discriminate between K+ and Rb+, while R224A and R228A had reduced inhibition of short-circuit current for both ions, when analyzed by voltage clamping of M. sexta midguts.  相似文献   

13.
After the transfer of prototype plasmids R6K (IncX), R387 (IncK), R27 (IncH1) and T (IncN) to E. coli M nalr the appearance of histidine-dependent mutants (R27, T), histidine-leucine-dependent mutants (R6K), methionine-proline-dependent mutants (R387) was observed among the resulting transconjugates. The mutations of E. coli M nalr R+ cells induced by the introduction of the plasmids were accompanied by the transformation of the cells from the S-form into the R-form. In contrast to the prototrophs E. coli M nalr, the auxotrophs carrying plasmids R6K, R27, T acquired sensitivity to phage T7, and the methionine-proline-dependent mutant became sensitive to phages T and T7. The above-mentioned plasmids rendered E. coli M cells capable of synthetizing the donor pili. But the adsorption of phages T3 and T7 on the auxotrophic cells, both with and without plasmids, occurred due to their interaction with the cell-wall receptors.  相似文献   

14.
The lipid extract of the temperate red alga Rhodymenia pertusa has yielded four eicosanoid metabolites, three of which are new natural products. Using principally NMR and MS techniques, their structures were deduced as 5R,6S-dihydroxy-7(E),9(E),11(Z),14(Z)-eicosatetraenoic acid (5R,6S-diHETE), 5R*,6S*-dihydroxy-7(E),9(E),11(Z),14(Z),17(Z)-eicosapentaenoic acid (5R*,6S*-diHEPE), 5-hydroxy-6(E),8(Z),11(Z),14(Z)-eicosatetraenoic acid (5-HETE), 5-hydroxy-6(E),8(Z),11(Z),14(Z),17(Z)-eicosapentaenoic acid (5-HEPE). The co-occurrence of these metabolites strongly suggests that R. pertusa contains a unique 5R-lipoxygenase system acting on both arachidonic and eicosapentaenoic acids.  相似文献   

15.
A membrane protein complex, succinate dehydrogenase (SQR) from Escherichia coli has been purified and crystallised. This enzyme is composed of four subunits containing FAD, three iron-sulphur clusters and one haem b as prosthetic groups. The obtained crystals belong to the hexagonal space group P6(3) with the unit-cell dimensions of a=b=123.8 A and c=214.6 A. An asymmetric unit of the crystals contains one SQR monomer (M(r) 120 kDa). A data set is now available at 4.0 A resolution with 88.1% completeness and 0.106 R(merge). We have obtained a molecular replacement solution that shows sensible molecular packing, using the soluble domain of E. coli QFR (fumarate reductase) as a search model. The packing suggests that E. coli SQR is a crystallographic trimer rather than a dimer as observed for the E. coli QFR.  相似文献   

16.
This report is devoted to study of the hydrocarbon composition of the extract of buds of birch (family Betulaceae). We identified 3,5-dihydroxy-7,4??-dimethoxyflavone, 5,7-dihydroxy-3,4??-dimethoxyflavone, 7-methoxy-4??,5-dihydroxyflavanone, 4??-methoxy-5,7-hydroxyflavanone, 5??,6??-dihydroxycaryophyllen-4(14),8(15)-diene ((1R,5R,6R,9S)-5,6-dihydroxy-11,11-dimethyl-4,8-methylenebicyclo[7.2.0]-undecane), 5??,6??-dihydroxycaryophyllen-3,8(15)-diene ((1R,5R,6R,9S)-5,6-dihydroxy-11,11-dimethyl-8-methylenebicyclo[7.2.0]undec-3-ene), 6??,9??-dihydroxy-??-humulene((1E,4E,6R,9S)-6,9-dihydroxy-4,11,11-trimethyl-8-methylene-1,4-cycloundecanediene), 5??,6??,8??-trihydroxycariolan, and (1R,5R,6R,8R,9S)-trihydroxycariolan. The gas chromatographic retention indices of all identified compounds were determined.  相似文献   

17.
We have developed a novel method for constructing synthetic genes that encode a series of peptides comprising perfect repeat motifs based on a high molecular weight subunit (HMW glutenin subunit), a highly repetitive storage protein from wheat seed. A series of these genes of sequentially increasing size was produced, four of which (called R3, 4, 5, 6) were expressed in Escherichia coli. Activity of the synthetic genes in E. coli was confirmed by Northern blot analysis but SDS-PAGE of crude protein extracts failed to show any expressed peptides when stained using Coomassie brilliant blue R250. However, Western blots probed with a HMW glutenin subunit-specific polyclonal antibody showed the presence of the R6 peptide (M(r) 22005) in the crude cell extracts and both this and the R3 peptide (M(r) 12005) were subsequently purified by extraction with hot aqueous ethanol followed by precipitation with acetone and separated by RP-HPLC. The R4 and R5 peptides were not purified. The purified R3 and R6 peptides absorbed Coomassie brilliant blue R250 or other protein stains only weakly and this was considered to account for their failure to be revealed by staining of separations of the crude protein extracts. Circular dichroism spectroscopy showed that both peptides had similar beta-turn rich structures similar to the repetitive sequences present in the whole HMW glutenin subunits. We conclude that expression of perfect repeat peptides in E. coli is a suitable system for the study of structure-function relationships in wheat gluten proteins and other highly repetitive proteins.  相似文献   

18.
Halohydrin dehalogenase from Agrobacterium radiobacter AD1 (HheC) is a valuable tool in the preparation of R enantiomers of epoxides and β-substituted alcohols. In contrast, the halohydrin dehalogenase from Arthrobacter sp. AD2 (HheA) shows a low S enantioselectivity toward most aromatic substrates. Here, three amino acids (V136, L141, and N178) located in the two neighboring active-site loops of HheA were proposed to be the key residues for controlling enantioselectivity. They were subjected to saturation mutagenesis aimed at evolving an S-selective enzyme. This led to the selection of two outstanding mutants (the V136Y/L141G and N178A mutants). The double mutant displayed an inverted enantioselectivity (from S enantioselectivity [E(S)] = 1.7 to R enantioselectivity [E(R)] = 13) toward 2-chloro-1-phenylethanol without compromising enzyme activity. Strikingly, the N178A mutant showed a large enantioselectivity improvement (E(S) > 200) and a 5- to 6-fold-enhanced specific activity toward (S)-2-chloro-1-phenylethanol. Further analysis revealed that those mutations produced some interference for the binding of nonfavored enantiomers which could account for the observed enantioselectivities. Our work demonstrated that those three active-site residues are indeed crucial in modulating the enantioselectivity of HheA and that a semirational design strategy has great potential for rapid creation of novel industrial biocatalysts.  相似文献   

19.
Arginine257 (R257), in the de-helix that caps the Q(B) site of the D1 protein, has been shown by mutational studies to play a key role in the sensitivity of Photosystem II (PS II) to bicarbonate-reversible binding of the formate anion. In this article, the role of this residue has been further investigated through D1 mutations (R257E, R257Q, and R257K) in Chlamydomonas reinhardtii. We have investigated the activity of the Q(B) site by studying differences from wild type on the steady-state turnover of PS II, as assayed through chlorophyll (Chl) a fluorescence yield decay after flash excitation. The effects of p-benzoquinone (BQ, which oxidizes reduced Q(B), Q(B)(-) ) and 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU, which blocks electron flow from Q(A)(-) to Q(B)) were measured. The equilibrium constants of the two-electron gate were obtained through thermoluminescence measurements. The thermoluminescence properties were changed in the mutants, especially when observed after pretreatment with 100 microM BQ. A theoretical analysis of the thermoluminescence data, based mainly on the recombination pathways model of Rappaport et al. (2005), led to the conclusion that the free-energy difference for the recombination of Q(B)(-) with S(2) was reduced by 20-40 mV in the three mutants (D1-R257K, D1-R257Q, and D1-R257E); this was interpreted to be due to a lowering of the redox potential of Q(B)/Q(B)(-). Further, since the recombination of Q(A)(-) with S(2) was unaffected, we suggest that no significant change in redox potential of Q(A)/Q(A)(-) occurred in these three mutants. The maximum variable Chl a fluorescence yield is lowered in the mutants, in the order R257K > R257Q > R257E, compared to wild type. Our analysis of the binary oscillations in Chl a fluorescence following pretreatment of cells with BQ showed that turnover of the Q(B) site was relatively unaffected in the three mutants. The mutant D1-R257E had the lowest growth rate and steady-state activity and showed the weakest binary oscillations. We conclude that the size and the charge of the amino acid at the position D1-257 play a role in PS II function by modulating the effective redox potential of the Q(B)/Q(B)(-) pair. We discuss an indirect mechanism mediated through electrostatic and/or surface charge effects and the possibility of more pleiotropic effects arising from decreased stability of the D1/D2 and D1/CP47 interfaces.  相似文献   

20.
Respiration of stems and branches of trees (R(S)) has typically been estimated by measuring radial CO(2) efflux from woody tissue (E(A)) and rates of efflux are often scaled temporally using a temperature relationship (Q(10)). High concentrations of CO(2) in xylem sap ([CO(2)*]) have been shown to affect E(A), and the transport of CO(2) in the xylem stream has been suggested as a mechanism to explain field observations of temperature-independent fluctuations in E(A). Sap velocity and temperature were manipulated in detached branch segments of sycamore (Platanus occidentalis L.) under controlled conditions to quantify these effects. Within individual branches of similar size, E(A) and [CO(2)*] were greater at low sap velocity, while the amount of respired CO(2) transported in sap (transport flux, F(T)) was greater at high sap velocity. E(A) was linearly correlated with [CO(2)*]. In branches of three diameter classes (1, 2, and 3 cm), volume-based E(A), F(T), and R(S) did not differ, but surface-area based CO(2) fluxes increased with diameter class. Regardless of diameter, E(A) accounted for only 30% of respired CO(2) at high sap velocity, while at low sap velocity, E(A) accounted for 71% of respired CO(2). E(A), F(T), and R(S) measured at 5, 20, and 35 degrees C at the same sap velocity showed a typical exponential response to temperature. However, at the lowest temperature, E(A) accounted for only 18% of the CO(2) released from respiring cells compared with 44% at the highest temperature, perhaps due to the effect of temperature on the solubility of CO(2) in water. These results directly demonstrate the transport of respired CO(2) in the xylem stream and may help to explain inconsistencies in stem and branch respiration measurements made in situ.  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