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 60 毫秒
1.
Summary Using a direct Monte Carlo simulation, population growth of helper T-cells (N H) and viral cells (N v) is studied for an immune response model with an enhanced spatial inter-cellular interaction relevant to HIV as a function of viral mutation. In the absence of cellular mobility (P mob=0), the helper T-cells grow nonmonotonically before reaching saturation and the viral population grows monotonically before reaching a constant equilibrium. Cellular mobility (P mob=1) enhances the viral growth and reduces the stimulative T-cell growth. Below a mutation threshold (P c), the steady-state density of helper T-cell (p H) is larger than that of the Virus (p v); the density difference Δp o(=pV−pH) remains a constant at P mob=1 while −Δp o→0 as P mutP c at P mob=0. Above the mutation threshold, the difference Δp o in cell density, grows with ΔP=P mutP c monotonically: ΔP o ∞ (ΔP)β ≃ with β≈0.574±0.016 in absence of mobility, while Δp o≈6(ΔP) with P mob=1.  相似文献   

2.
To determine the photosynthetic characteristics of C3 plants and their sensitivity to CO2 at different altitudes on the Tibetan Plateau, hulless barley (Hordeum vulgare L. ssp. vulgare) was grown at altitudes of 4,333 m and 3,688 m. Using gas-exchange measurements, photosynthetic parameters were simulated, including the maximum net photosynthesis (P max) and the apparent quantum efficiency (α). Plants growing at higher altitude had higher net photosynthetic rates (P N), photosynthesis parameters (P max and α) and sensitivities to CO2 enhancement than plants growing at lower altitude on the Tibetan Plateau. The enhancements of P N, P max, and α for plants growing at higher altitude, corresponding with 10 μmol(CO2) mol−1 increments, were approximately 0.20∼0.45%, 0.05∼0.20% and 0.12∼0.36% greater, respectively, than for plants growing at lower altitude, respectively, where CO2 levels rose from 10 to 170 μmol(CO2) mol−1. Therefore, on the Tibetan Plateau, the changes in the photosynthetic capacities and the photosynthetic sensitivities to CO2 observed in the C3 plants grown above 3,688 m are likely to increase with altitude despite the decreasing CO2 partial pressure.  相似文献   

3.
The linkage of pair-wise interactions of contact site mutations of HbS has been studied using Le Lamentin [His-20 (α)→Gln], Hoshida [Glu-43 (β)→Gln] and α2β2T87Q mutations as the prototype of three distinct classes of contact sites of deoxy HbS fiber. Binary mixture experiments established that βA-chain with the Thr-87 (β)→Gln mutation is as potent as the γ-chain of HbF (α2γ2) in inhibiting polymerization. On combining the influence of Le Lamentin mutation with that of β2T87Q mutations; the net influence is only partial additivity. On the other hand, in binary mixture studies, combined influence of Hoshida mutation with that of β2T87Q mutations is synergistic. Besides, a significant level of synergistic complementation is also seen when the Le Lamentin and Hoshida mutations are combined in HbS (symmetrical tetramers). Le Lamentin and Hoshida mutation introduced into the cis-dimer of the asymmetric hybrid tetramer completely neutralizes the Val-6 (β) dependent polymerization. Accordingly, we propose that combining the perturbation of intra-double strand contact site with that of an inter-double strand contact site exhibit synergy when they are present in two different chains of the αβ dimer. A comparison of the present results with that of the earlier studies suggest that when the two contact site perturbations are from the same sub-unit of the αβ dimer only partial additivity is observed. The map of interaction linkage of the contact site mutations exposes new strategies in the design of novel anti-sickling Hbs for the gene therapy of sickle cell disease.  相似文献   

4.
Killing of wild-type spores of Bacillus subtilis by t-butyl hydroperoxide, cumene hydroperoxide and peracetic acid was not through DNA damage, as shown by the absence of mutations in the survivors and the identical sensitivity of spores of strains with or without a recA mutation. In contrast, B. subtilis spores (termed αβ) lacking the DNA protective α/β-type small, acid-soluble spore proteins (SASP) were more sensitive to t-butyl hydroperoxide and cumene hydroperoxide, and their killing was in large part through DNA damage, as shown by the high frequency of mutations in the survivors and the greater sensitivity of αβ recA spores. Analysis of t-butyl hydroperoxide-treated spores showed that generation of DNA damage in αβ spores was more rapid than in wild-type spores; α/β-type SASP also protected against DNA strand breakage in vitro caused by t-butyl hydroperoxide. α/β-Type SASP appeared to play no role in protection of spores from killing by peracetic acid; wild-type and αβ spores exhibited identical peracetic acid sensitivity and their killing by this agent appeared to be not through DNA damage. Received 17 December 1996/ Accepted in revised form 13 March 1997  相似文献   

5.
The α1 subunit coding for the human brain type E calcium channel (Schneider et al., 1994) was expressed in Xenopus oocytes in the absence, and in combination with auxiliary α2δ and β subunits. α1E channels directed with the expression of Ba2+ whole-cell currents that completely inactivated after a 2-sec membrane pulse. Coexpression of α1E with α2bδ shifted the peak current by +10 mV but had no significant effect on whole-cell current inactivation. Coexpression of α1E with β2a shifted the peak current relationship by −10 mV, and strongly reduced Ba2+ current inactivation. This slower rate of inactivation explains that a sizable fraction (40 ± 10%, n= 8) of the Ba2+ current failed to inactivate completely after a 5-sec prepulse. Coinjection with both the cardiac/brain β2a and the neuronal α2bδ subunits increased by ≈10-fold whole-cell Ba2+ currents although coinjection with either β2a or α2bδ alone failed to significantly increase α1E peak currents. Coexpression with β2a and α2bδ yielded Ba2+ currents with inactivation kinetics similar to the β2a induced currents, indicating that the neuronal α2bδ subunit has little effect on α1E inactivation kinetics. The subunit specificity of the changes in current properties were analyzed for all four β subunit genes. The slower inactivation was unique to α1E2a currents. Coexpression with β1a, β1b, β3, and β4, yielded faster-inactivating Ba2+ currents than currents recorded from the α1E subunit alone. Furthermore, α1E2bδ/β1a; α1E2bδ/β1b; α1E2bδ/β3; α1E2bδ/β4 channels elicited whole-cell currents with steady-state inactivation curves shifted in the hyperpolarized direction. The β subunit-induced changes in the properties of α1E channel were comparable to modulation effects reported for α1C and α1A channels with β3≈β1b > β1a≈β4≫β2a inducing fastest to slowest rate of whole-cell inactivation. Received: 27 March 1997/Revised: 10 July 1997  相似文献   

6.
Ten plant species belonging to 5 families,i.e., Cruciferae, Cucurbitaceae, Gramineae, Leguminosae, and Solanaceae, were grown in a sand soil at two pH levels. The soil was subjected to an application of CdCl2 at rates of 0 to 700 mgCd·kg−1 soil. The relationship between Cd concentration in the shoots (tc) and soil (sc; NH4NO3 extractable) was expressed by the equation: log (tc)=α+βlog (sc). The coefficients α and β were estimated for each species at each level of soil pH. Plottings of the scores on α and β axes showed that the Cd accumulation characteristics in the plants appeared to depend on the families irrespective of soil pH. Based on theupper critical concentration of Cd in the tops (Ct), the Cruciferae and Leguminosae species were found to be the most and the least tolerant to Cd, respectively. The Ct values correlated exceedingly well with the values of α.  相似文献   

7.
Four temperature treatments were studied in the climate controlled growth chambers of the Georgia Envirotron: 25/20, 30/25, 35/30, and 40/35 °C during 14/10 h light/dark cycle. For the first growth stage (V3-5), the highest net photosynthetic rate (P N) of sweet corn was found for the lowest temperature of 28–34 μmol m−2 s−1 while the P N for the highest temperature treatment was 50–60 % lower. We detected a gradual decline of about 1 P N unit per 1 °C increase in temperature. Maximum transpiration rate (E) fluctuated between 0.36 and 0.54 mm h−1 (≈5.0–6.5 mm d−1) for the high temperature treatment and the minimum E fluctuated between 0.25 and 0.36 mm h−1 (≈3.5–5.0 mm d−1) for the low temperature treatment. Cumulative CO2 fixation of the 40/35 °C treatment was 33.7 g m−2 d−1 and it increased by about 50 % as temperature declined. The corresponding water use efficiency (WUE) decreased from 14 to 5 g(CO2) kg−1(H2O) for the lowest and highest temperature treatments, respectively. Three main factors affected WUE, P N, and E of Zea: the high temperature which reduced P N, vapor pressure deficit (VPD) that was directly related to E but did not affect P N, and quasi stem conductance (QC) that was directly related to P N but did not affect E. As a result, WUE of the 25/20 °C temperature treatment was almost three times larger than that of 40/35 °C temperature treatment.  相似文献   

8.
Phylogenetic relationships among reptiles were examined using previously published and newly determined hemoglobin sequences. Trees reconstructed from these sequences using maximum-parsimony, neighbor-joining, and maximum-likelihood algorithms were compared with a phylogenetic tree of Amniota, which was assembled on the basis of published morphological data. All analyses differentiated α chains into αA and αD types, which are present in all reptiles except crocodiles, where only αA chains are expressed. The occurrence of the αD chain in squamates (lizards and snakes only in this study) appears to be a general characteristic of these species. Lizards and snakes also express two types of β chains (βI and βII), while only one type of β chain is present in birds and crocodiles. Reconstructed hemoglobin trees for both α and β sequences did not yield the monophyletic Archosauria (i.e., crocodilians + birds) and Lepidosauria (i.e., Sphenodon+ squamates) groups defined by the morphology tree. This discrepancy, as well as some other poorly resolved nodes, might be due to substantial heterogeneity in evolutionary rates among single hemoglobin lineages. Estimation of branch lengths based on uncorrected amino acid substitutions and on distances corrected for multiple substitutions (PAM distances) revealed that relative rates for squamate αA and αD chains and crocodilian β chains are at least twice as high as those of the rest of the chains considered. In contrast to these rate inequalities between reptilian orders, little variation was found within squamates, which allowed determination of absolute evolutionary rates for this subset of hemoglobins. Rate estimates for hemoglobins of lizards and snakes yielded 1.7 (αA) and 3.3 (β) million years/PAM when calibrated with published divergence time vs. PAM distance correlates for several speciation events within snakes and for the squamate ↔ sphenodontid split. This suggests that hemoglobin chains of squamate reptiles evolved ∼3.5 (αA) or ∼1.7 times (β) faster than their mammalian equivalents. These data also were used to obtain a first estimate of some intrasquamate divergence times. Received: 15 September 1997 / Accepted: 4 February 1998  相似文献   

9.
 Mutant mice generated by disrupting the H2-Aa b major histocompatibility complex (Mhc) gene are demonstrated here to express Aβb chains in the absence of α chains. These mice possess a CD4+ helper T cell (Th) repertoire that uses predominantly the Vβ7 T-cell antigen receptor (Tcr) segment for recognition of any protein antigen presented by the α-free Aβ molecule. As an alloantigen, the Aα-free Aβ molecule is recognized very poorly by T cells from a series of class II disparate mouse strains, indicating that it is grossly different from normal α/β heterodimers. Indeed, molecular modeling suggests a β/β homodimer arrangement with an altered geometry of the Tcr contact area. Interestingly, the mutant mice exhibit normal alloreactivity, without a restricted Vβ usage, toward a series of foreign α/β class II heterodimers, although their T cells developed in the absence of such heterodimers. Thus, the complementarity of Tcr to normal α/β heterodimers, and thereby also alloreactivity, appears to be an ontogeny independent (i. e., germline-encoded) feature. Received: 30 September 1996 / Revised: 18 October 1996  相似文献   

10.
The Na,K-ATPases and the H,K-ATPases are two potassium-dependent homologous heterodimeric P2-type pumps that catalyze active transport of Na+ in exchange for K+ (Na,K-ATPase) or H+ in exchange for K+ (H,K-ATPase). The ubiquitous Na,K-ATPase maintains intracellular ion balance and membrane potential. The gastric H,K-ATPase is responsible for acid secretion by the parietal cell of the stomach. Both pumps consist of a catalytic α-subunit and a glycosylated β-subunit that is obligatory for normal pump maturation and trafficking. Individual N-glycans linked to the β-subunits of the Na,K-ATPase and H,K-ATPase are important for stable membrane integration of their respective α subunits, folding, stability, subunit assembly, and enzymatic activity of the pumps. They are also essential for the quality control of unassembled β-subunits that results in either the exit of the subunits from the ER or their ER retention and subsequent degradation. Overall, the importance of N-glycans for the␣maturation and quality control of the H,K-ATPase is greater than that of the Na,K-ATPase. The roles of individual N-glycans of the β-subunits in the post-ER trafficking, membrane targeting and plasma membrane retention of the Na,K-ATPase and H,K-ATPase are different. The Na,K-ATPase β 1-subunit is the major β-subunit isoform in cells with lateral location of the pump. All three N-glycans of the Na,K-ATPase β 1-subunit are important for the lateral membrane retention of the pump due to glycan-mediated interaction between the β 1-subunits of the two neighboring cells in the cell monolayer and cytosolic linkage of the α-subunit to the cytoskeleton. This intercellular β 1β 1 interaction is also important for formation of cell–cell contacts. In contrast, the N-glycans unique to the Na,K-ATPase β 2-subunit,which has up to eight N-glycosylation sites, contain apical sorting information. This is consistent with the apical location of the Na,K-ATPase in normal and malignant epithelial cells with high abundance of the β 2-subunit. Similarly, all seven N-glycans of the gastric H,K-ATPase β-subunit determine apical sorting of this subunit. Supported in part by NIH grants DK46917, DK58333, D53642, and USVA  相似文献   

11.
A method for microperfusion of isolated segments of the midgut epithelium of Drosophila larvae has been developed to characterize cellular transport pathways and membrane transporters. Stereological ultrastructural morphometry shows that this epithelium has unusually long tight junctions, with little or no lateral intercellular volume normally found in most epithelia. Amplification of the apical and basal aspects of the cells, by ≈ 17-fold and ≈ 7-fold, respectively, predicts an almost exclusively transcellular transport system for solutes. This correlates with the high lumen-negative transepithelial potential (Vt) of 38 to 45 mV and high resistance (Rt) of 800 to 1400 Ω • cm2 measured by terminated cable analysis, in contrast to other microperfused epithelia like the renal proximal tubule. Several blockers (amiloride 10−4 M, ouabain 10−4 M, bumetanide 10−4 M), K+-free solutions, or organic solutes such as D-glucose 10 mM or DL-alanine 0.5 mM failed to affect Vt or Rt. Bafilomycin-A1 (3 to 5 μM) decreased Vt by ≈ 40% and short-circuit current (Isc) by ≈ 50%, and decreased intracellular pH when applied from the basal side only, consistent with an inhibition of an electrogenic V-H+-ATPase located in the basal membrane. Gradients of H+ were detected by pH microelectrodes close to the basal aspect of the cells or within the basal extracellular labyrinth. The apical membrane is more conductive than the basal membrane, facilitating secretion of base (presumably HCO3), driven by the basal V-H+-ATPase.  相似文献   

12.
Summary α,β-Dehydroamino acids are useful peptide modifiers. However, their stereoelectronic properties still remain insufficiently recognized. Based on FTIR experiments in the range ofv s(N-H), AI, AII andv s(Cα=Cβ) and ab initio calculations with B3LYP/6–31G*, we studied the solution conformational preferences and the amide electron density perturbation of Ac-ΔXaa-NHMe, where ΔXaa=ΔAla, (E)-ΔAbu, (Z)-ΔAbu, (Z)-ΔLeu, (Z)-ΔPhe and ΔVal. Each of these dehydroamides adopts a C5 structure, which in Ac-ΔAla-NHMe is fully extended and accompanied by the strong C5 hydrogen bond. Interaction with bond Cα=Cβ lessens the amidic resonance within the flanking amide groups. TheN-terminal C=O bond is noticeably shorter, both amide bonds are longer than the corresponding bonds in the saturated entities and the N-terminal amide system is distorted. Ac-ΔAla-NHMe constitutes an exception. ItsC-terminal amide bond is shorter than the standard one and both amide systems are ideally planar. Ac-(E)-ΔAbu-NHMe shares stereoelectronic features with both Ac-ΔAla-NHMe and (Z)-dehydroamides.  相似文献   

13.
Cochlear outer hair cells (OHCs) terminally differentiate prior to the onset of hearing. During this time period, thyroid hormone (TH) dramatically influences inner ear development. It has been shown recently that TH enhances the expression of the motor protein prestin via liganded TH receptor β (TRβ) while in contrast the expression of the potassium channel KCNQ4 is repressed by unliganded TRα1. These different mechanisms of TH regulation by TRα1 or TRβ prompted us to analyse other ion channels that are required for the final differentiation of OHCs. We analysed the onset of expression of the Ca2+ channel CaV1.3, and the K+ channels SK2 and BK and correlated the results with the regulation via TRα1 or TRβ. The data support the hypothesis that proteins expressed in rodents prior to or briefly after birth like CaV1.3 and prestin are either independent of TH (e.g. CaV1.3) or enhanced through TRβ (e.g. prestin). In contrast, proteins expressed in rodents later than P6 like KCNQ4 (∼P6), SK2 (∼P9) and BK (∼P11) are repressed through TRα1. We hypothesise that the precise regulation of expression of the latter genes requires a critical local TH level to overcome the TRα1 repression. Harald Winter and Claudia Braig contributed equally to this work.  相似文献   

14.
The first synthesis of the Neu5Gc analogue of SiaT n disaccharide, which can be detected in breast tumors by immunochemical methods, is reported. The regioselective sialylation of (3-trifluoroacetamidopropyl)-2-azido-2-deoxy-α-D-galactopyranoside with peracetate of the methyl ester ofN-acetoxyacetyl-neuraminic acid β-ethylthioglycoside in the presence ofN-iodosuccinimide and trifluoromethanesulfonic acid (or its trimethylsilyl ester) resulted in the derivatives of α- and β-sialyl(2→6)galactosaminide in 39 and 32% yields, respectively. The catalytic hydrogenolysis of the azido group and subsequentN- andO-acetylation of the α-anomer gave the peracetate of trifluoroacetamidopropyl glycoside. Removal of the protective groups led to glycoside Neu5Gcα2→6GalNAcα-O(CH2)3NH2. Using the Neu5Gc derivative with acetoxyacetyl groups at positions O9 and O4 as a donor increases the α-selectivity of sialylation to afford the α- and β-anomers in 69 and 8% yields, respectively.  相似文献   

15.
Small-angle neutron scattering (SANS) on the unilamellar vesicle (ULV) populations (diameter 500 and 1,000 Å) in D2O was used to characterize lipid vesicles from dimyristoylphosphatidylcholine (DMPC) at three phases: gel Lβ′, ripple Pβ′ and liquid Lα. Parameters of vesicle populations and internal structure of the DMPC bilayer were characterized on the basis of the separated form factor (SFF) model. Vesicle shape changes from nearly spherical in the Lα phase to elliptical in the Pβ′ and Lβ′ phases. This is true for vesicles prepared via extrusion through pores with the diameter 500 Å. Parameters of the internal bilayer structure (thickness of the membrane and the hydrophobic core, hydration and the surface area of the lipid molecule) were determined on the basis of the hydrophobic–hydrophilic (HH) approximation of neutron scattering length density across the bilayer ρ(x) and of the step function (SF) approximation of ρ(x). DMPC membrane thickness in the Lα phase (T=30°C) demonstrates a dependence on the membrane curvature for extruded vesicles. Prepared via extrusion through 500 Å diameter pores, vesicle population in the Lα phase has the following characteristics: average value of minor semi-axis 266±2 Å, ellipse eccentricity 1.11±0.02, polydispersity 26%, thickness of the membrane 48.9±0.2 Å and of the hydrophobic core 19.9±0.4 Å, surface area 60.7±0.5 Å2 and number of water molecules 12.8±0.3 per DMPC molecule. Vesicles prepared via extrusion through pores with the diameter 1,000 Å have polydispersity of 48% and membrane thickness of 45.5±0.6 Å in the Lα phase. SF approximation was used to describe the DMPC membrane structure in Lβ′ (T=10°C) and Pβ′ (T=20°C) phases. Extruded DMPC vesicles in D2O have membrane thickness of 49.6±0.5 Å in the Lβ′ phase and 48.3±0.6 Å in the Pβ′ phase. The dependence of the DMPC membrane thickness on temperature was restored from the SANS experiment.  相似文献   

16.
Despite the progress in studies of the properties and functions of low-threshold calcium channels (LTCCs) [1], the mechanisms of their selectivity and permeability remain unstudied in detail. We performed a comparative analysis of the selectivity of three cloned pore-forming LTCC subunits (α1G, α1H, and α1I) functionally expressed in Xenopus oocytes with respect to bivalent alkaline-earth metal cations (Ba2+, Ca2+, and Sr2+. The relative conductivities (G) of these channels were determined according to the amplitudes of macroscopic currents (I) and potentials of zero currents (E). The currents were recorded after preliminary intracellular injection of a fast calcium buffer, BAPTA, in order to suppress the endogenous calcium-dependent chloride conductivity. Channels formed by α1G subunits demonstrated the following ratios of the amplitudes of macroscopic currents and potentials of zero current: I Ca:I Ba:I Sr = 1.00:0.75:1.12 and E CaE BaE Sr. For channels that were formed by α1H and α1I subunits, these ratios were as follows: I Ca:I Ba:I Sr = 1.00:1.20:1.17, E CaE BaE Sr and I Ca:I Ba:I Sr = 1.00:1.48: 1.45, E CaE BaE Sr respectively. The different macroscopic conductivities and similar potentials of zero current typical of α1G and α1I channels indicate that, probably, various bivalent cations can in a differential manner influence the stochastic parameters of functioning of these channels. At the same time, channels formed by α1H subunits are characterized by more positive potentials of zero current for Ca2+. It seems possible that the selectivity of the above channels is determined by mechanisms that mediate the selectivity of most high-threshold calcium channels (more affine binding of Ca2+ inside the pore). Neirofiziologiya/Neurophysiology, Vol. 37, No. 4, pp. 319–329, July–August, 2005.  相似文献   

17.
The polyphasic patterns of fluorescence induction rise in pea leaves in vivo and after the treatment with ionophores have been studied using a Plant Efficiency Analyzer. To analyze in detail photosystem II (PS II) electron transfer processes, an extended PS II model was applied, which included the sums of exponential functions to specify explicitly the light-driven formation of the transmembrane electric potential (ΔΨ(t)) as well as pH in the lumen (pHL(t)) and stroma (pHS(t)). PS II model parameters and numerical coefficients in ΔΨ(t), pHL(t), and pHS(t) were evaluated to fit fluorescence induction data for different experimental conditions: leaf in vivo or after ionophore treatment at low or high light intensity. The model imitated changes in the pattern of fluorescence induction rise due to the elimination of transmembrane potential in the presence of ionophores, when ΔΨ = 0 and pHL(t), pHS(t) changed to small extent relative to control values in vivo, with maximum ΔΨ(t) ∼ 90 mV and ΔΨ(t) ∼ 40 mV for the stationary state at ΔpH ≅ 1.8. As the light intensity was increased from 300 to 1200 μmol m−2 s−1, the heat dissipation rate constants increased threefold for nonradiative recombination of P680+Phe and by ∼30% for P680+QA. The parameters ΔΨ, pHS and pHL were analyzed as factors of PS II redox state populations and fluorescence yield. The kinetic mechanism of fluorescence quenching is discussed, which is related with light-induced lumen acidification, when +QA and P680+ recombination probability increases to regulate the QA reduction.  相似文献   

18.
PCNA is an essential factor for DNA replication, repair, chromatin metabolism, and effector of cell-cycle regulatory signals. The assignment of backbone 1HN, 13Cα, 13CO, and 15N, and sidechain 13Cβ resonances of the human PCNA homotrimeric ring (∼90 kDa, 261 residues) is reported here.  相似文献   

19.
Coherent processes in an initial phase of charge transfer in reaction centers (RCs) of the triple mutant S(L178)K/G(M203)D/L(M214)H of Rhodobacter sphaeroides were investigated by difference (light — dark) absorption spectroscopy with 18 fsec time resolution. Electron transfer in the B cofactor branch is activated in this mutant, while the A-branch electron transfer is slowed in comparison with native RCs of Rba. sphaeroides. A bulk of absorption difference spectra was analyzed in the 940–1060 nm range (stimulated emission of excited bacteriochlorophyll dimer P* and absorption of bacteriochlorophyll anions BA and β, where β is a bacteriochlorophyll substituting the native bacteriopheophytin HA) and in the 735–775 nm range (bleaching of the absorption band of the bacteriopheophytin HB in the B-branch) in the −0.1 to 4 psec range of delays with respect to the moment of photoexcitation of P at 870 nm. Spectra were measured at 293 and 90 K. The kinetics of P* stimulated emission at 940 nm shows its decay with a time constant of ∼14 psec at 90 K and ∼18 psec at 293 K, which is accompanied by oscillations with a frequency of ∼150 cm−1. A weak absorption band is found at 1018 nm that is formed ∼100 fsec after excitation of P and reflects the electron transfer from P* to β and/or BA with accumulation of the P+β and/or P+BA states. The kinetics of ΔA at 1018 nm contains the oscillations at ∼150 cm−1 and distinct low-frequency oscillations at 20–100 cm−1; also, the amplitude of the oscillations at 150 cm−1 is much smaller at 293 than at 90 K. The oscillations in the kinetics of the 1018 nm band do not contain a 32 cm−1 mode that is characteristic for native Rba. sphaeroides RCs having water molecule HOH55 in their structure. The ΔA kinetics at 751 nm reflects the electron transfer to HB with formation of the P+HB state. The oscillatory part of this kinetics has the form of a single peak with a maximum at ∼50 fsec completely decaying at ∼200 fsec, which might reflect a reversible electron transfer to the B-branch. The results are analyzed in terms of coherent nuclear wave packet motion induced in the P* excited state by femtosecond light pulses, of an influence of the incorporated mutations on the mutual position of the energy levels of charge separated states, and of the role of water HOH55 in the dynamics of the initial electron transfer.  相似文献   

20.
A series of N α-acyl (alkyl)- and N α-alkoxycarbonyl-derivatives of l- and d-ornithine were prepared, characterized, and analyzed for their potency toward the bacterial enzyme N α-acetyl-l-ornithine deacetylase (ArgE). ArgE catalyzes the conversion of N α-acetyl-l-ornithine to l-ornithine in the fifth step of the biosynthetic pathway for arginine, a necessary step for bacterial growth. Most of the compounds tested provided IC50 values in the μM range toward ArgE, indicating that they are moderately strong inhibitors. N α-chloroacetyl-l-ornithine (1g) was the best inhibitor tested toward ArgE providing an IC50 value of 85 μM while N α-trifluoroacetyl-l-ornithine (1f), N α-ethoxycarbonyl-l-ornithine (2b), and N α-acetyl-d-ornithine (1a) weakly inhibited ArgE activity providing IC50 values between 200 and 410 μM. Weak inhibitory potency toward Bacillus subtilis-168 for N α-acetyl-d-ornithine (1a) and N α-fluoro- (1f), N α-chloro- (1g), N α-dichloro- (1h), and N α-trichloroacetyl-ornithine (1i) was also observed. These data correlate well with the IC50 values determined for ArgE, suggesting that these compounds might be capable of getting across the cell membrane and that ArgE is likely the bacterial enzymatic target.  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