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.

Introduction

The Elongator complex, comprising six subunits (Elp1p-Elp6p), is required for formation of 5-carbamoylmethyl (ncm5) and 5-methoxycarbonylmethyl (mcm5) side chains on wobble uridines in 11 out of 42 tRNA species in Saccharomyces cerevisiae. Loss of these side chains reduces the efficiency of tRNA decoding during translation, resulting in pleiotropic phenotypes. Overexpression of hypomodified \( {\text {tRNA}_{{\rm s^{2} {\rm UUU}}}^{{\rm Lys}} , {\rm tRNA}_{{\rm s^{2} {\rm UUG}}}^{{\rm Gln }} \;{\rm and}\;{\rm tRNA}_{{\rm s^{2} {\rm UUC}}}^{{\rm Glu}}} \), which in wild-type strains are modified with mcm5s2U, partially suppress phenotypes of an elp3Δ strain.

Objectives

To identify metabolic alterations in an elp3Δ strain and elucidate whether these metabolic alterations are suppressed by overexpression of hypomodified \( {\text {tRNA}_{{\rm s^{2} {\rm UUU}}}^{{\rm Lys}} , {\rm tRNA}_{{\rm s^{2} {\rm UUG}}}^{{\rm Gln }} \;{\rm and}\;{\rm tRNA}_{{\rm s^{2} {\rm UUC}}}^{{\rm Glu}}} \).

Method

Metabolic profiles were obtained using untargeted GC-TOF-MS of a temperature-sensitive elp3Δ strain carrying either an empty low-copy vector, an empty high-copy vector, a low-copy vector harboring the wild-type ELP3 gene, or a high-copy vector overexpressing \( {\text {tRNA}_{{\rm s^{2} {\rm UUU}}}^{{\rm Lys}} , {\rm tRNA}_{{\rm s^{2} {\rm UUG}}}^{{\rm Gln }} \;{\rm and}\;{\rm tRNA}_{{\rm s^{2} {\rm UUC}}}^{{\rm Glu}}} \). The temperature sensitive elp3Δ strain derivatives were cultivated at permissive (30 °C) or semi-permissive (34 °C) growth conditions.

Results

Culturing an elp3Δ strain at 30 or 34 °C resulted in altered metabolism of 36 and 46 %, respectively, of all metabolites detected when compared to an elp3Δ strain carrying the wild-type ELP3 gene. Overexpression of hypomodified \( {\text {tRNA}_{{\rm s^{2} {\rm UUU}}}^{{\rm Lys}} , {\rm tRNA}_{{\rm s^{2} {\rm UUG}}}^{{\rm Gln }} \;{\rm and}\;{\rm tRNA}_{{\rm s^{2} {\rm UUC}}}^{{\rm Glu}}} \) suppressed a subset of the metabolic alterations observed in the elp3Δ strain.

Conclusion

Our results suggest that the presence of ncm5- and mcm5-side chains on wobble uridines in tRNA are important for metabolic homeostasis.
  相似文献   

2.
3.
The acidity constants of twofold protonated, antivirally active, acyclic nucleoside phosphonates (ANPs), H(2) (PE)(±) , where PE(2-) =9-[2-(phosphonomethoxy)ethyl]adenine (PMEA(2-) ), 2-amino-9-[2-(phosphonomethoxy)ethyl]purine (PME2AP(2-) ), 2,6-diamino-9-[2-(phosphonomethoxy)ethyl]purine (PMEDAP(2-) ), or 2-amino-6-(dimethylamino)-9-[2-(phosphonomethoxy)ethyl]purine (PME(2A6DMAP)(2-) ), as well as the stability constants of the corresponding ternary Cu(Arm)(H;PE)(+) and Cu(Arm)(PE) complexes, where Arm=2,2'-bipyridine (bpy) or 1,10-phenanthroline (phen), are compared. The constants for the systems containing PE(2-) =PMEDAP(2-) and PME(2A6DMAP)(2-) have been determined now by potentiometric pH titrations in aqueous solution at I=0.1M (NaNO(3) ) and 25°; the corresponding results for the other ANPs were taken from our earlier work. The basicity of the terminal phosphonate group is very similar for all the ANP(2-) species, whereas the addition of a second amino substituent at the pyrimidine ring of the purine moiety significantly increases the basicity of the N(1) site. Detailed stability-constant comparisons reveal that, in the monoprotonated ternary Cu(Arm)(H;PE)(+) complexes, the proton is at the phosphonate group, that the ether O-atom of the ?CH(2) ?O?CH(2) ?P(O)$\rm{{_{2}^{-}}}$(OH) residue participates, next to the P(O)$\rm{{_{2}^{-}}}$(OH) group, to some extent in Cu(Arm)(2+) coordination, and that π?π stacking between the aromatic rings of Cu(Arm)(2+) and the purine moiety is rather important, especially for the H?PMEDAP(-) and H?PME(2A6DMAP)(-) ligands. There are indications that ternary Cu(Arm)(2+) -bridged stacks as well as unbridged (binary) stacks are formed. The ternary Cu(Arm)(PE) complexes are considerably more stable than the corresponding Cu(Arm)(R?PO(3) ) species, where R?PO$\rm{{_{3}^{2-}}}$ represents a phosph(on)ate ligand with a group R that is unable to participate in any kind of intramolecular interaction within the complexes. The observed stability enhancements are mainly attributed to intramolecular-stack formation in the Cu(Arm)(PE) complexes and also, to a smaller extent, to the formation of five-membered chelates involving the ether O-atom present in the ?CH(2) ?O?CH(2) ?PO$\rm{{_{3}^{2-}}}$ residue of the PE(2-) species. The quantitative analysis of the intramolecular equilibria involving three structurally different Cu(Arm)(PE) isomers shows that, e.g., ca. 1.5% of the Cu(phen)(PMEDAP) system exist with Cu(phen)(2+) solely coordinated to the phosphonate group, 4.5% as a five-membered chelate involving the ether O-atom of the ?CH(2) ?O?CH(2) ?PO$\rm{{_{3}^{2-}}}$ residue, and 94% with an intramolecular π?π stack between the purine moiety of PMEDAP(2-) and the aromatic rings of phen. Comparison of the various formation degrees of the species formed reveals that, in the Cu(phen)(PE) complexes, intramolecular-stack formation is more pronounced than in the Cu(bpy)(PE) species. Within a given Cu(Arm)(2+) series the stacking intensity increases in the order PME2AP(2-) 相似文献   

4.
Assimilatory nitrite reductase (aNiR) reduces nitrite to an ammonium ion and has siroheme and a [Fe(4)S(4)] cluster as prosthetic groups. A reaction mechanism for Nii3, an aNiR from tobacco, is proposed based on high resolution X-ray structures and UV-Vis (ultraviolet-visible) microspectroscopy of Nii3-ligand complexes. Analysis of UV-Vis spectral changes in Nii3 crystals with increasing X-ray exposure showed prosthetic group reductions. In Nii3-NO2(-) structures, X-ray irradiation enhanced the progress of the reduction reaction, and cleavage of the N-O bond was observed when X-ray doses were increased. Crystal structures of Nii3 with other bound ligands, such as Nii3-NO and Nii3-NH(2)OH, were also determined. Further, by combining information from these Nii3 ligand-bound structures, including that of Nii3-NO2(-), with UV-Vis microspectral data obtained using different X-ray doses, a reaction mechanism for aNiR was suggested. Cleavage of the two N-O bonds of nitrite was envisaged as a two-step process: first, the N-O bond close to Lys224 was cleaved, followed by cleavage of the N-O bond close to Arg109. X-ray structures also indicated that aNiR-catalyzed nitrite reduction proceeded without the need for conformation changes in active site residues. Geometrical changes in the ligand molecules and the placement of neighboring water molecules appeared to be important to the stability of the active site residue interactions (Arg109, Arg179, and Lys224) and the ligand molecule. These interactions may contribute to the efficiency of aNiR reduction reactions.  相似文献   

5.
Deletion of the substituted pyrimidine ring in purine-2'-deoxynucleoside 5'-monophosphates leads to the artificial nucleotide analog dImMP(2-) . This analog can be incorporated into DNA to yield, upon addition of Ag(+) ions, a molecular wire. Here, we measured the acidity constants of H(2) (dImMP)(±) having one proton at N(3) and one at the PO$\rm{{_{3}^{2-}}}$ group by potentiometric pH titrations in aqueous solution. The micro acidity constants show that N(3) is somewhat more basic than PO$\rm{{_{3}^{2-}}}$ and, consequently, the (H??dImMP)(-) tautomer with the proton at N(3) dominates to ca. 75%. The calculated micro acidity constants are confirmed by (31) P- and (1) H-NMR chemical shifts. The assembled data allow many quantitative comparisons, e.g., the N(3)-protonated and thus positively charged imidazole residue facilitates deprotonation of the P(O)(2) (OH)(-) group by 0.3?pK units. Information on the intrinsic site basicities also allows predictions about metal-ion binding; e.g., Mg(2+) and Mn(2+) will primarily coordinate to the phosphate group, whereas Ni(2+) and Cu(2+) will preferably bind to N(3). Macrochelate formation for these metal ions is also predicted. The micro acidity constant for N(3)H(+) deprotonation in the (H???dImMP?H)(±) species (pk(a) 6.46) and the M(n+) -binding properties are of relevance for understanding the behavior of dImMP units present in DNA hairpins and metalated duplexes.  相似文献   

6.
R Glasser  E J Gabbay 《Biopolymers》1968,6(2):243-254
The synthesis of spermine derivatives (II), \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm R}_1 {\rm R}_{\rm 2} {\rm R}_{\rm 3} \mathop {\rm N}\limits^ + \left( {{\rm CH}_2 } \right)_3 \mathop {\rm N}\limits^ + {\rm R}_{\rm 1} {\rm R}_{\rm 2} \left( {{\rm CH}_2 } \right)_2 ]_2 \cdot 4{\rm X}^ - $\end{document}, and spermidine derivatives (III), \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm R}_1 {\rm R}_{\rm 2} {\rm R}_{\rm 3} \mathop {\rm N}\limits^ + \left( {{\rm CH}_2 } \right)_4 \mathop {\rm N}\limits^ + {\rm R}_{\rm 1} {\rm R}_{\rm 2} \left( {{\rm CH}_2 } \right)_3 \mathop {\rm N}\limits^ + {\rm R}_{\rm 1} {\rm R}_{\rm 2} {\rm R}_3 \cdot 3{\rm X}^ - $\end{document}, are reported. The effects of these salts on the helix–coil transition of rA–rU and rI–rC helices were examined. Increasing the size of the hydrophobic substituents, R1, R2, and R3 lowers the degree of stabilization of the helical structure. The disproportionation reaction, 2rA–rU→rA–rU2 + rA occurs readily with salts II and III, especially when the substituents, R1, R2, and R3 are small, i.e., H or Me. Spermine is found to stabilize the rA–rU2 and rI–rC helices to approximately the same extent; however, large differences between the degree of stabilization of rA–rU2 and rI-rC helices are observed when the substituents R1, R2, and R3 are large hydrophobic groups. Similar results are also obtained for the spermidine series. Finally, differences in the interactions of the salts II and III with rA–rU2 and rI–rC helices suggest that the latter helix is denser.  相似文献   

7.
Arsenite (As(III)) is the predominant arsenic (As) species in reducing environments. As(III) is less strongly adsorbed than As(V) at circumneutral pH conditions by common non‐iron metal oxides in sediments such as those of aluminum. Therefore, oxidation of As(III) to As(V) could contribute to an improved immobilization of As and thus help mitigate As contamination in groundwater. Microbial oxidation of As(III) is known to readily under aerobic conditions, however, the dissolved oxygen (O2) concentration in groundwater may be limited due to the poor solubility of O2 and its high chemical reactivity with reduced compounds. Nitrate (${\rm NO}_{3}^{{-} } $ ), can be considered as an alternative electron acceptor, which can support oxidation of As(III) to As(V) by denitrifying bacteria. In this study, two up‐flow sediment columns packed with activated alumina (AA) were utilized to demonstrate the role of denitrification on the oxidation of As(III) to As(V) and its contribution to improved As adsorption onto AA. One column was supplied with ${\rm NO}_{3}^{{-} } $ (C1) and its performance was compared with a control column lacking ${\rm NO}_{3}^{{-} } $ (C2). During most of the operation when the pH was in the circumneutral range (days 50–250), the release of arsenic was greater from C2 compared to C1. The effluent As concentrations started increasing on days 60 and 100 in C2 and C1, respectively. Complete breakthrough started on day 200 in C2; whereas in C1, complete breakthrough was never achieved. The effluent and solid phase As speciation was dominated by As(V) in C1, indicating the occurrence of As(III) oxidation due to ${\rm NO}_{3}^{{-} } $ ; whereas in C2, only As(III) was dominant. This study illustrates a bioremediation or natural attenuation process based on anoxic microbial ${\rm NO}_{3}^{{-} } $ ‐dependent oxidation of As(III) to more readily adsorbed As(V) as a means to enhance the immobilization of As on alumina oxide particles in subsurface environments. Biotechnol. Bioeng. 2010;107: 786–794. © 2010 Wiley Periodicals, Inc.  相似文献   

8.
Nitrogen (N) retention by tree canopies is believed to be an important process for tree nutrient uptake, and its quantification is a key issue in determining the impact of atmospheric N deposition on forest ecosystems. Due to dry deposition and retention by other canopy elements, the actual uptake and assimilation by the tree canopy is often obscured in throughfall studies. In this study, 15N-labeled solutions ( $ ^{15} {\text{NH}}_{4}^{ + } $ and $ ^{15} {\text{NO}}_{3}^{ - } $ ) were used to assess dissolved inorganic N retention by leaves/needles and twigs of European beech, pedunculate oak, silver birch, and Scots pine saplings. The effects of N form, tree species, leaf phenology, and applied $ {\text{NO}}_{3}^{ - } $ to $ {\text{NH}}_{4}^{ + } $ ratio on the N retention were assessed. Retention patterns were mainly determined by foliar uptake, except for Scots pine. In twigs, a small but significant 15N enrichment was detected for $ {\text{NH}}_{4}^{ + } $ , which was found to be mainly due to physicochemical adsorption to the woody plant surface. The mean $ {{^{15} {\text{NH}}_{4}^{ + } } \mathord{\left/ {\vphantom {{^{15} {\text{NH}}_{4}^{ + } } {^{15} {\text{NO}}_{3}^{ - } }}} \right. \kern-0em} {^{15} {\text{NO}}_{3}^{ - } }} $ retention ratio varied considerably among species and phenological stadia, which indicates that the use of a fixed ratio in the canopy budget model could lead to an over- or underestimation of the total N retention. In addition, throughfall water under each branch was collected and analyzed for $ ^{15} {\text{NH}}_{4}^{ + } $ , $ ^{15} {\text{NO}}_{3}^{ - } $ , and all major ions. Net throughfall of $ ^{15} {\text{NH}}_{4}^{ + } $ was, on average, 20 times higher than the actual retention of $ ^{15} {\text{NH}}_{4}^{ + } $ by the plant material. This difference in $ ^{15} {\text{NH}}_{4}^{ + } $ retention could not be attributed to pools and fluxes measured in this study. The retention of $ ^{15} {\text{NH}}_{4}^{ + } $ was correlated with the net throughfall of K+, Mg2+, Ca2+, and weak acids during leaf development and the fully leafed period, while no significant relationships were found for $ ^{15} {\text{NO}}_{3}^{ - } $ retention. This suggests that the main driving factors for $ {\text{NH}}_{4}^{ + } $ retention might be ion exchange processes during the start and middle of the growing season and passive diffusion at leaf senescence. Actual assimilation or abiotic uptake of N through leaves and twigs was small in this study, for example, 1–5% of the applied dissolved 15N, indicating that the impact of canopy N retention from wet deposition on forest productivity and carbon sequestration is likely limited.  相似文献   

9.
This work reviews the brown-rot fungal biochemical mechanism involved in the biodegradation of lignified plant cell walls. This mechanism has been acquired as an apparent alternative to the energetically expensive apparatus of lignocellulose breakdown employed by white-rot fungi. The mechanism relies, at least in the incipient stage of decay, on the oxidative cleavage of glycosidic bonds in cellulose and hemicellulose and the oxidative modification and arrangement of lignin upon attack by highly destructive oxygen reactive species such as the hydroxyl radical generated non-enzymatically via Fenton chemistry \( ({\text{F}}{{\text{e}}^{{{3} + }}} + {{\text{H}}_{{2}}}{{\text{O}}_{{2}}} \to {\text{F}}{{\text{e}}^{{{2} + }}} + \cdot {\text{OH}}{{ + }^{ - }}{\text{OH}}) \). Modifications in the lignocellulose macrocomponents associated with this non-enzymatic attack are believed to aid in the selective, near-complete removal of polysaccharides by an incomplete cellulase suite and without causing substantial lignin removal. Utilization of this process could provide the key to making the production of biofuel and renewable chemicals from lignocellulose biomass more cost-effective and energy efficient. This review highlights the unique features of the brown-rot fungal non-enzymatic, mediated Fenton reaction mechanism, the modifications to the major plant cell wall macrocomponents, and the implications and opportunities for biomass processing for biofuels and chemicals.  相似文献   

10.
A novel benzoylphloroglucinol derivative, garcimultiflorone G ( 1 ), was isolated from the fruits of Garcinia multiflora. The structure of 1 was determined through extensive 1D‐ and 2D‐NMR, and MS analyses. Garcimultiflorone G ( 1 ) showed inhibitory effects against superoxide anion (O$\rm{{_{2}^{{^\cdot} -}}}$ ) generation and elastase release by human neutrophils in response to formyl‐L ‐methionyl‐L ‐leucyl‐L ‐phenylalanine/cytochalasin B (fMLP/CB), with IC50 values of 6.97±1.56 and 11.70±1.58 μM , respectively.  相似文献   

11.
The pathway and ab initio direct kinetics of the decomposition 5-aminotetrazole (5-ATZ) to HN3 and NH2CN was investigated. Reactant, products and transition state were optimized with MP2 and B3LYP methods using 6–311G** and aug-cc-pVDZ basis sets. The intrinsic reaction coordinate curve of the reaction was calculated using the MP2 method with 6–311G** basis set. The energies were refined using CCSD(T)/6–311G**. Rate constants were evaluated by conventional transition-state theory (CVT) and canonical variational transition-state theory (TST), with tunneling effect over 300 to 2,500 K. The results indicated that the tunneling effect and the variational effect are small for the calculated rate constants. The fitted three-parameter expression calculated using the CVT and TST methods are and , respectively. Figure The mechanism of the decomposition process of 5-ATZ to HN3 and NH2CN  相似文献   

12.
13.
A consecutive, first-order, irreversible, biochemical reaction, \documentclass{article}\pagestyle{empty}\begin{document}$ A{\textstyle{{k(\theta)} \over {{\rm Enzyme }1}}} \to B{\textstyle{{k(\theta)} \over {{\rm Enzyme 2}}}} \to C $\end{document}, taking place in a series of N reactors with product recycle is considered. A discrete version of the maximum principle is used to derive general equations necessary for maximizing the production of (1) the final product, C, by choosing the temperature or the pH value in each reactor, and (2) the intermediate product, B, by choosing the reactor volume. A numerical computation for a series of three reactors with recycle is illustrated. The effects of varying the recycle rates on the optimal state and decision variables are also presented.  相似文献   

14.
Methanobacterium thermoautotrophicum was grown in continuous culture in a fermenter gassed with H2 and CO2 as sole carbon and energy sources, and in a medium which contained either NH4Cl or gaseous N2 as nitrogen source. Growth was possible with N2. Steady states were obtained at various gas flow rates with NH4Cl and with and the maintenance coefficient varied with the gas input and with the nitrogen source. Growth of Methanococcus thermolithotrophicus in continuous culture in a fermenter gassed with H2, CO2 as nitrogen, carbon and energy sources was also examined.Abbreviations molecular growth yield (g dry weight of cells per mol of CH4 evolved) - growth rate (h-1) - D dilution rate (h-1) - rate (h-1); relation of Neijssel and Tempest and of Stouthamer and Bettenhaussen - energy  相似文献   

15.
Kinetics of ethanol inhibition in alcohol fermentation   总被引:3,自引:0,他引:3  
The inhibitory effect of ethanol on yeast growth and fermentation has been studied for the strain Saccharomyces cerevisiae ATCC No. 4126 under anaerobic batch conditions. The results obtained reveal that there is no striking difference between the response of growth and ethanol fermentation. Two kinetic models are also proposed to describe the kinetic pattern of ethanol inhibition on the specific rates of growth and ethanol fermentation: \documentclass{article}\pagestyle{empty}\begin{document}$$\begin{array}{*{20}c} {\frac{{\mu _i }}{{\mu _0 }} = 1{\rm } - {\rm }\left( {\frac{P}{{P_m }}} \right);\alpha } \hfill & {\left( {{\rm for}\ {\rm growth}} \right)} \hfill \\ {\frac{{\nu _i }}{{\nu _0 }} = 1{\rm } - {\rm }\left( {\frac{P}{{P'_m }}} \right);\beta } \hfill & {\left( {{\rm for}\ {\rm ethanol}\ {\rm production}} \right)} \hfill \\ \end{array}$$\end{document} The maximum allowable ethanol concentration above which cells do not grow was predicted to be 112 g/L. The ethanol-producing capability of the cells was completely inhibited at 115 g/L ethanol. The proposed models appear to accurately represent the experimental data obtained in this study and the literature data.  相似文献   

16.
Azotochelin is a biscatecholate siderophore produced by the nitrogen-fixing soil bacterium Azotobacter vinelandii. The complexation properties of azotochelin with a series of oxoanions [Mo(VI), W(VI) and V(V)] and divalent cations [Cu(II), Zn(II), Co(II) and Mn(II)] were investigated by potentiometry, UV–vis and X-ray spectroscopy. Azotochelin forms a strong 1:1 complex with molybdate (log K = 7.6 ± 0.4) and with tungstate and vanadate; the stability of the complexes increases in the order Mo < V < W (log K appMo = 7.3 ± 0.4; log K appV = 8.8 ± 0.4 and log K appW = 9.0 ± 0.4 at pH 6.6). The Mo atom in the 1:1 Mo–azotochelin complex is bound to two oxo groups in a cis position and to the two catecholate groups of azotochelin, resulting in a slightly distorted octahedral configuration. Below pH 5, azotochelin appears to form polynuclear complexes with Mo in addition to the 1:1 complex. Azotochelin also forms strong complexes with divalent metals. Of the metals studied, Cu(II) binds most strongly to azotochelin , followed by Zn(II) , Mn(II) and Co(II) . Since very few organic ligands are known to bind strongly to oxoanions (and particularly molybdate) at circumneutral pH, the unusual properties of azotochelin may be used for the separation and concentration of oxoanions in the laboratory and in the field. In addition, azotochelin may prove useful for the investigation of the biogeochemistry of Mo, W and V in aquatic and terrestrial systems. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users.  相似文献   

17.
The data processing method of the turbidimetric bioassay of nisin was modified to facilitate its industrial application. The influence of the initial indicator concentration was minimized by a redefined specific dose of the bacteriocin as the quotient between the titer of the added bacteriocin and the initial population density of the indicator in the suspension. It was found that d c = 0.125 μg ml−1 was the critical dose of nisin that can cause a complete inhibition of the indicator, Pediococcus acidilactici UL5, with an initial OD of 0.135. To eliminate the interference of the cell debris, an equation, , exploiting d c, was formulated to obtain the intrinsic survival proportion. The use of the specific dose of the bacteriocin and the intrinsic survival proportion as parameters of the dose/response curve greatly enhanced its repeatability and feasibility. A dual-dosage approach was developed to further simplify the conventional standard dose/response curve method.  相似文献   

18.
The threshold for rotation about the yaw axis was determined for constant acceleration stimuli as a function of their duration in the range from 3 to 25 s. From the torsion-swing model the following theoretical equation can be derived: 1 $$a_{{\text{thr}}} = {C \mathord{\left/ {\vphantom {C {\left[ {1 - \exp \left( { - {{t_s } \mathord{\left/ {\vphantom {{t_s } {\tau _1 }}} \right. \kern-\nulldelimiterspace} {\tau _1 }}} \right)} \right]}}} \right. \kern-\nulldelimiterspace} {\left[ {1 - \exp \left( { - {{t_s } \mathord{\left/ {\vphantom {{t_s } {\tau _1 }}} \right. \kern-\nulldelimiterspace} {\tau _1 }}} \right)} \right]}}$$ , where a thr=acceleration amplitude at threshold, t s =duration of the acceleration, τ1=time constant, C=threshold for very long stimuli. According to this formula the Mulder product (i.e. the product of the threshold acceleration amplitude and the duration of the stimulus) is constant for durations up to 0.3 τ1. The best fit of this theoretical function to the somatosensory data is found for τ1=14.5 s, and C=0.220/s 2. The time within the Mulder product is constant (about 5s) is doubtless due to the mechanics of the semicircular canals. For the oculogyral data a lower value of τ1 is found. We do not have any explanation for this lower value.  相似文献   

19.
Tobacco expresses four isomers of assimilatory nitrite reductase (aNiR), leaf‐type (Nii1 and Nii3), and root‐type (Nii2 and Nii4). The high‐resolution crystal structures of Nii3 and Nii4, determined at 1.25 and 2.3 Å resolutions, respectively, revealed that both proteins had very similar structures. The Nii3 structure provided detailed geometries for the [4Fe–4S] cluster and the siroheme prosthetic groups. We have generated two types of Nii3 variants: one set focuses on residue Met175 (Nii3‐M175G, Nii3‐M175E, and Nii3‐M175K), a residue that is located on the substrate entrance pathway; the second set targets residue Gln448 (Nii3‐Q448K), a residue near the prosthetic groups. Comparison of the structures and kinetics of the Nii3 wild‐type (Nii3‐WT) and the Met175 variants showed that the hydrophobic side‐chain of Met175 facilitated enzyme efficiency (kcat/Km). The Nii4‐WT has Lys449 at the equivalent position of Gln448 in Nii3‐WT. The enzyme activity assay revealed that the turnover number (kcat) and Michaelis constant (Km) of Nii4‐WT were lower than those of Nii3‐WT. However, the kcat/Km of Nii4‐WT was about 1.4 times higher than that of Nii3‐WT. A comparison of the kinetics of the Nii3‐Q448K and Nii4‐K449Q variants revealed that the change in kcat/Km was brought about by the difference in Residue 448 (defined as Gln448 in Nii3 and Lys449 in Nii4). By combining detailed crystal structures with enzyme kinetics, we have proposed that Nii3 is the low‐affinity and Nii4 is the high‐affinity aNiR.  相似文献   

20.
This work discloses two bona fide gates through which the CO ligand can leave the distal cavity of carboxy human cytoglobin, reaching the solvent. The investigation was based on molecular dynamics, aided by a minimal randomly‐oriented force applied to the ligand. The shortest pathway progresses toward the main gate, H81‐R84, in the open state, with the H81 imidazole moiety turned toward the solvent. A longer pathway develops toward the diametrically opposed W31‐W151 gate. In between, CO may be entrapped into binding cavities, either along the path toward the gates, or in a cul‐de‐sac, from which CO may even be incapable to escape. This behavior contrasts with carboxy myoglobin, where the corresponding H64 gate, when opened, is the sole used by CO to get to the solvent. These observations, which could hold also for other small ligands of biological interest, such as O2, NO, and NO$\rm{{_{3}^{-}}}$ , provide an answer to a neglected aspect of the mysterious six‐coordinated globins.  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