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.
1. Membrane particles prepared from ultrasonically-disrupted, aerobically-grown Escherichia coli were centrifuged on to a plastic film that was supported perpendicular to the centrifugal field to yield oriented membrane multilayers. In such preparations, there is a high degree of orientation of the planes of the membranes such that they lie parallel to each other and to the supporting film. 2. When dithionite- or succinate-reduced multilayers are rotated in the magnetic field of an e.p.r. spectrometer, about an axis lying in the membrane plane, angular-dependent signals from an iron-sulphur cluster at g(x)=1.92, g(y)=1.93 and g(z)=2.02 are seen. The g=1.93 signal has maximal amplitude when the plane of the multilayer is perpendicular to the magnetic field. Conversely, the g=2.02 signal is maximal when the plane of the multilayer is parallel with the magnetic field. 3. Computer simulations of the experimental data show that the cluster lies in the cytoplasmic membrane with the g(y) axis perpendicular to the membrane plane and with the g(x) and g(z) axes lying in the membrane plane. 4. In partially-oxidized multilayers, a signal resembling the mitochondrial high-potential iron-sulphur protein (Hipip) is seen whose g(z)=2.02 axis may be deduced as lying perpendicular to the membrane plane. 5. Appropriate choice of sample temperature and receiver gain reveals two further signals in partially-reduced multilayers: a g=2.09 signal arises from a cluster with its g(z) axis in the membrane plane, whereas a g=2.04 signal is from a cluster with the g(z) axis lying along the membrane normal. 6. Membrane particles from a glucose-grown, haem-deficient mutant contain dramatically-lowered levels of cytochromes and exhibit, in addition to the iron-sulphur clusters seen in the parental strain, a major signal at g=1.90. 7. Only the latter may be demonstrated to be oriented in multilayer preparations from the mutant. 8. Comparisons are drawn between the orientations of the iron-sulphur proteins in the cytoplasmic membrane of E. coli and those in mitochondrial membranes. The effects of diminished cytochrome content on the properties of the iron-sulphur proteins are discussed.  相似文献   

2.
Appropriate values for the orientation of the amide transition dipoles are essential to the growing use of isotopically edited vibrational spectroscopy generally in structural biology and to infrared dichroism measurements on membrane-associated alpha-helices, in particular. The orientations of the transition moments for the amide vibrations of an alpha-helix have been determined from the ratio of intensities of the A- and E(1)-symmetry modes in the infrared spectra of poly(gamma-methyl-L-glutamate)(x)-co-(gamma-n-octadecyl-L-glutamate)( y) oriented on silicon substrates. Samples possessing a high degree of alignment were used to facilitate band fitting. Consistent results were obtained from both attenuated total reflection and transmission experiments with polarized radiation, yielding values of Theta(I) = 38 degrees, Theta(II) = 73 degrees, and Theta(A) = 29 degrees, relative to the helix axis, for the amide I, amide II, and amide A bands, respectively. The measurements are discussed both in the context of the somewhat divergent older determinations, and in relation to the helix geometry and results on model amide compounds, to resolve current uncertainties in the literature.  相似文献   

3.
Metabolic and ventilatory variables were measured in a large semifossorial marsupial, the hairy-nosed wombat (Lasiorhinus latifrons, 21.9 kg). In normoxia, the rate of oxygen consumption was 63% of that predicted for a similar-sized marsupial, and the level of ventilation (V(E)) was such that the convective requirement (V(E)/VO2) was similar to other mammals. Exposure to hypercapnia (5% CO(2)) evoked a hyperventilatory response (3.55 x normoxia) that was no different to that observed for epigeal (surface-dwelling) marsupials; the increase in V(E) was primarily achieved with an increase in tidal volume. Exposure to hypoxia (15% to 8% O(2)) resulted in a hyperventilation (principally through an increase in frequency), although the response was blunted (in 8% O(2), 1.85 x normoxia) and only at the severest levels did hypometabolism contribute. The attenuated response to hypoxia in the wombat is presumably a reflection of a semifossorial lifestyle and a tolerance to this respiratory stimulant.  相似文献   

4.
Ca(2+) and Cl(-) ions are essential elements for the oxygen evolution activity of photosystem II (PSII). It has been demonstrated that these ions can be exchanged with Sr(2+) and Br(-), respectively, and that these ion exchanges modify the kinetics of some electron transfer reactions at the Mn?Ca cluster level (Ishida et al., J. Biol. Chem. 283 (2008) 13330-13340). It has been proposed from thermoluminescence experiments that the kinetic effects arise, at least in part, from a decrease in the free energy level of the Mn(4)Ca cluster in the S? state though some changes on the acceptor side were also observed. Therefore, in the present work, by using thin-layer cell spectroelectrochemistry, the effects of the Ca(2+)/Sr(2+) and Cl(-)/Br(-) exchanges on the redox potential of the primary quinone electron acceptor Q(A), E(m)(Q(A)/Q(A)(-)), were investigated. Since the previous studies on the Ca(2+)/Sr(2+) and Cl(-)/Br(-) exchanges were performed in PsbA3-containing PSII purified from the thermophilic cyanobacterium Thermosynechococcus elongatus, we first investigated the influences of the PsbA1/PsbA3 exchange on E(m)(Q(A)/Q(A)(-)). Here we show that i) the E(m)(Q(A)/Q(A)(-)) was up-shifted by ca. +38mV in PsbA3-PSII when compared to PsbA1-PSII and ii) the Ca(2+)/Sr(2+) exchange up-shifted the E(m)(Q(A)/Q(A)(-)) by ca. +27mV, whereas the Cl(-)/Br(-) exchange hardly influenced E(m)(Q(A)/Q(A)(-)). On the basis of the results of E(m)(Q(A)/Q(A)(-)) together with previous thermoluminescence measurements, the ion-exchange effects on the energetics in PSII are discussed.  相似文献   

5.
Water oxidation at photosystem II Mn-cluster is mediated by the redox-active tyrosine Y(Z). We calculated the redox potential (E(m)) of Y(Z) and its symmetrical counterpart Y(D), by solving the linearized Poisson-Boltzmann equation. The calculated E(m)(Y( )/Y(-)) were +926 mV/+694 mV for Y(Z)/Y(D) with the Mn-cluster in S2 state. Together with the asymmetric position of the Mn-cluster relative to Y(Z/D), differences in H-bond network between Y(Z) (Y(Z)/D1-His(190)/D1-Asn(298)) and Y(D) (Y(D)/D2-His(189)/D2-Arg(294)/CP47-Glu(364)) are crucial for E(m)(Y(Z/D)). When D1-His(190) is protonated, corresponding to a thermally activated state, the calculated E(m)(Y(Z)) was +1216 mV, which is as high as the E(m) for P(D1/D2). We observed deprotonation at CP43-Arg(357) upon S-state transition, which may suggest its involvement in the proton exit pathway. E(m)(Y(D)) was affected by formation of P(D2)(+) (but not P(D1)(+)) and sensitive to the protonation state of D2-Arg(180). This points to an electrostatic link between Y(D) and P(D2).  相似文献   

6.
The yields of nitrate and nitrite from decomposition of peroxynitrite in phosphate buffer at 37 degrees C were determined in the pH range 1-14. The NO(2)(-)/NO(3)(-) yields showed a stepwise variation with pH, with inflection points at approximately pH 3.1, 5.8, 6.8, 8.0, and 11.9. Nitrite formation increased strongly above pH 7 at the expense of nitrate, but above pH 12 nitrate again became the major product (80% at pH 14). At this pH, the Arrhenius parameters were E(a)=24.1+/-0.2kcal mol(-1) and A=(4.9+/-1.3)x10(12)s(-1). The yields of NO(2)(-), NO(3)(-), and O(2) measured at pH 5.8, 7.4, and 8.5 as a function of the initial peroxynitrite concentration (50-1000 microM) were linear only at pH 5.8. In the presence of carbon dioxide, oxygen production at pH 7.5 and pH 10 was found to be linear on the CO(2) concentration. The experimental observations were satisfactorily reproduced by kinetic simulations including principal component analyses. These data strongly suggest that the chemistry of peroxynitrite is exclusively mediated by z.rad;NO(2) and HO(z.rad;) radicals in the absence, and by z.rad;NO(2) and CO(3)(z.rad;-) radicals in the presence of CO(2).  相似文献   

7.
Conformational changes taking place in the gastric H+,K+-ATPase when shifting from the K+-induced E2 form to the E1 form upon replacing K+ ions by Na+ were investigated by different spectroscopic approaches. No significant secondary-structure change or secondary-structure reorientation with respect to the membrane plane could be measured by attenuated total reflection Fourier transform infrared spectroscopy of oriented films. Circular dichroism and Raman spectra obtained on tubulovesicle suspensions indicated no significant secondary structure or tyrosine and tryptophan side-chain environment changes in tubulovesicle suspensions. The smallest observable structural changes are discussed in term of the number of amino-acid residues involved for each technique.  相似文献   

8.
9.
The energy coupled NADH-ubiquinone (Q) oxidoreductase segment of the respiratory chain of Escherichia coli GR19N has been studied by EPR spectroscopy. Previously Matsushita et al. [(1987) Biochemistry 26, 7732-7737] have demonstrated the presence of two distinct NADH-Q oxidoreductases in E. coli membrane particles and designated them NADH dh I and NADH dh II. Although both enzymes oxidize NADH, only NADH dh I is coupled to the formation of the H+ electrochemical gradient. In addition to NADH, NADH dh I oxidizes nicotinamide hypoxanthine dinucleotide (deamino-NADH), while NADH dh II does not. In membrane particles we have detected EPR signals arising from four low-potential iron-sulfur clusters, one binuclear, one tetranuclear, and two fast spin relaxing g perpendicular = 1.94 type clusters (whose cluster structure has not yet been assigned). The binuclear cluster, temporarily designated [N-1]E, shows an EPR spectrum with gx,y,z = 1.92, 1.935, 2.03 and the Em7.4 value of -220 mV (n = 1). The tetranuclear cluster, [N-2]E, elicits a spectrum with gx,y,z = 1.90, 1.91, 2.05 and an Em7.4 of -240 mV (n = 1). These two clusters have been shown to be part of the NADH dh I complex by stability and inhibitor studies. When stored at 4 degrees C, both clusters are extremely labile as is the deamino-NADH-Q oxidoreductase activity. Addition of deamino-NADH in the presence of piericidin A results in nearly full reduction of [N-2]E within 17 s. In membrane particles pretreated with piericidin A, the cluster [N-1]E is only partly reducible by deamino-NADH and shows an altered line shape.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
A novel bis(6-ethylpicolinato)(H(2)O)oxovanadium(IV) complex (VO(6epa)(2) x (H(2)O)) was prepared and its structure was revealed by X-ray analysis (space group Pc(#7), a=10.838(2), b=11.148(5), c=16.642(3) A, and Z=2). Because VO(6epa)(2) x (H(2)O) exhibited higher in vitro insulinomimetic activity compared to that of vanadyl sulfate in terms of inhibition of free fatty acid (FFA) release from isolated rat adipocytes in the presence of epinephrine, its in vivo effect on whether the complex has a blood glucose normalizing effect was examined in KK-A(y) mice, a model animal of type 2 diabetes mellitus. VO(6epa)(2) x (H(2)O) was found to normalize the high blood glucose levels of KK-A(y) mice when given intraperitoneally at doses of 49 micromol/kg body weight for the first 4 days and then 39 micromol/kg body weight for 10 days. In addition, VO(6epa)(2) x (H(2)O) improved glucose tolerance ability as examined by the oral glucose test and seemed to have little toxicity in terms of serum parameters. VO(6epa)(2) x (H(2)O) showed higher normoglycemic activity than bis(6-methylpicolinato)oxovanadium(IV) (VO(6mpa)(2)) at the same dose. These results indicated that greater enhancement of the blood glucose normalizing effect in KK-A(y) mice by ethyl substitution compared to methyl substitution may be due to its being more strongly lipophilic.  相似文献   

11.
A procedure for the preparation of the fully reduced Cu(I) form of galactose oxidase, GOase(red), involving reduction of GOase(semi) (or GOase(ox)) with non-coordinating [Ru(NH(3))(6)](2+) (51 mV vs. nhe) is described. Air-free conditions and a two-fold excess of [Ru(NH(3))(6)](2+) give a stable product with no further UV-Vis changes over >1.5 h. Rate constants for the reduction of GOase(semi) (k(f)=860 M(-1) s(-1)) give a first-order [H(+)]-dependence (pK(1a)=7.9), but the reverse process involving [Ru(NH(3))(6)](3+) oxidation of GOase(red) (k(b)=18.6 M(-1) s(-1)) is independent of pH (5.5 to 9.5). The reduction potential E(2)(o)' (vs. nhe) for the GOase(semi)/GOase(red) (i.e. Cu(II)/Cu(I)) couple is 149 mV at pH 7.5, which varies from 160 mV (pH 5.5) to 120 mV (pH 10.5), suggesting pK(1a) (GOase(semi)) and pK(2a) (GOase(red)) acid dissociation constants both involving Tyr-495. It is concluded that pK(2a) is for acid dissociation of uncoordinated H(+)Tyr-495. Consistent with this interpretation rate constants/M(-1) s(-1) for the GOase(semi) Tyr495 Phe variant, k(f)=1.59x10(3) and k(b)=16.1, respectively, are independent of pH and give a reduction potential of 169 mV. Comparisons are made of reduction potentials (E(1)(o)'/mV pH 7.5) for the GOase(ox)/GOase(semi) (i.e. Tyr(.)/Tyr) couple, and are for the Cys228Gly variant (630), for enzyme with N(3)(-) for H(2)O at the substrate binding exogenous site (393), and for apo-protein (570). These compare with previously reported values for the variants Trp290His (730) and Tyr495Phe (450), and together serve to quantify different contributions to the unusually small E(1)(o)' of 400 mV for the Tyr(.)/Tyr couple. At pH 7.5 the reduction potential for the two-equivalent GOase(ox)/GOase(red) couple is calculated to be 275 mV. The rate constant for the reaction of GOase(red) with GOase(ox) is 4.4x10(3) M(-1) s(-1) at pH 7.5.  相似文献   

12.
D. Marsh 《Biophysical journal》1997,72(6):2710-2718
The transition moments for the amide bands from beta-sheet peptide structures generally do not exhibit axial symmetry about the director in linearly polarized Fourier transform infrared (FTIR) measurements on oriented systems. The angular dependences of the dichroic ratios of the amide bands are derived for beta-sheet structures in attenuated total reflection (ATR) and polarized transmission experiments on samples that are oriented with respect to the normal to the substrate and are randomly distributed with respect to the azimuthal angle in the plane of the orienting substrate. The orientational distributions of both the beta-strands and the beta-sheets are considered, and explicit expressions are given for the dichroic ratios of the amide I and amide II bands. The dichroic ratio of the amide II band, which is parallel polarized, can yield the orientation of the beta-strands directly, but to specify the orientations of the beta-sheets completely requires measurement of the dichroic ratios of both the amide I and amide II bands, or generally two bands with parallel and perpendicular polarizations. A random distribution in tilt of the planes of the beta-sheets does not give rise to equal dichroic ratios for bands with perpendicular and parallel polarizations, such as the amide I and amide II bands. The results are applied to previous ATR and polarized transmission FTIR measurements on a potassium channel-associated peptide, the Escherichia coli outer membrane protein OmpA, and the E. coli OmpF porin protein in oriented membranes.  相似文献   

13.
Mixture interactions between sour and salt taste modalities were investigated in rats by direct measurement of intracellular pH (pH(i)) and Na(+) activity ([Na(+)](i)) in polarized fungiform taste receptor cells (TRCs) and by chorda tympani (CT) nerve recordings. Stimulating the lingual surface with NaCl solutions adjusted to pHs ranging between 2.0 and 10.3 increased the magnitude of NaCl CT responses linearly with increasing external pH (pH(o)). At pH 7.0, the epithelial sodium channel (ENaC) blocker, benzamil, decreased NaCl CT responses and inhibited further changes in CT responses induced by varying pH(o) to 2.0 or 10.3. At constant pH(o), buffering NaCl solutions with potassium acetate/acetic acid (KA/AA) or HCO(3)(-)/CO(2) inhibited NaCl CT responses relative to CT responses obtained with NaCl solutions buffered with HEPES. The carbonic anhydrase blockers, MK-507 and MK-417, attenuated the inhibition of NaCl CT responses in HCO(3)(-)/CO(2) buffer, suggesting a regulatory role for pH(i). In polarized TRCs step changes in apical pH(o) from 10.3 to 2.0 induced a linear decrease in pH(i) that remained within the physiological range (slope = 0.035; r(2) = 0.98). At constant pH(o), perfusing the apical membrane with Ringer's solutions buffered with KA/AA or HCO(3)(-)/CO(2) decreased resting TRC pH(i), and MK-507 or MK-417 attenuated the decrease in pH(i) in TRCs perfused with HCO(3)(-)/CO(2) buffer. In parallel experiments, TRC [Na(+)](i) decreased with (a) a decrease in apical pH, (b) exposing the apical membrane to amiloride or benzamil, (c) removal of apical Na(+), and (d) acid loading the cells with NH(4)Cl or sodium acetate at constant pH(o). Diethylpyrocarbonate and Zn(2+), modification reagents for histidine residues in proteins, attenuated the CO(2)-induced inhibition of NaCl CT responses and the pH(i)-induced inhibition of apical Na(+) influx in TRCs. We conclude that TRC pH(i) regulates Na(+)-influx through amiloride-sensitive apical ENaCs and hence modulates NaCl CT responses in acid/salt mixtures.  相似文献   

14.
Fahmy K  Sakmar TP  Siebert F 《Biochemistry》2000,39(34):10607-10612
A highly conserved carboxylic acid residue in rhodopsin, Glu(134), modulates transducin (G(t)) interaction. It has been postulated that Glu(134) becomes protonated upon receptor activation. We studied the interaction between rhodopsin and G(t) using Fourier transform infrared (FTIR) difference spectroscopy combined with attenuated total reflection (ATR). Formation of the complex between G(t) and photoactivated rhodopsin reconstituted into phosphatidylcholine vesicles caused prominent infrared absorption increases at 1641, 1550, and 1517 cm(-)(1). The rhodopsin mutant E134Q was also studied. When measured in the presence of G(t), replacement of Glu(134) by glutamine abolished the low-frequency part of a broad absorption band at 1735 cm(-)(1) that is normally superimposed on the light-induced absorption changes of Asp(83) and Glu(122) of rhodopsin. In addition, a negative absorption band at 1400 cm(-)(1) that is evoked by interaction of native metarhodopsin II (MII) with G(t) was not observed in the difference spectrum of the E134Q mutant. Thus, Glu(134) is ionized in the dark and exhibits a symmetrical COO(-) stretching vibration at 1400 cm(-)(1). Glu(134) becomes protonated in the G(t)-MII complex and displays a C=O stretching mode near 1730 cm(-)(1). The E134Q mutation also affects absorption changes attributable to lipids, suggesting that the protonation of Glu(134) is linked to transfer of the carboxylic acid side chain from a polar to a nonpolar environment by becoming exposed to the lipid phase when G(t) binds. These results show directly that Glu(134) becomes protonated in MII upon G(t) binding and suggest that changes in receptor conformation affect lipid-protein interactions.  相似文献   

15.
Synthesis of fully hydrophilic star-shaped macromolecules with different kinds of arms (A(x)B(y)C(z)) based on polyglycidol (PGL, A(x)) and poly(ethylene oxide) (PEO, C(z)) arms and diepoxy compounds (diglycidyl ethers of ethylene glycol (DGEG) or neopentyl glycol (DGNG) in the core, B(y)) forming the core is described. Precursors of arms were prepared by polymerization of glycidol with protected -OH groups. The first-generation stars were formed in the series of consecutive-parallel reactions of arms A(x) with diepoxy compounds (B). These first-generation stars (A(x)B(y)), having approximately O-, Mt+ groups on the cores, were used as multianionic initiators for the second generation of arms (C(z)) built by polymerization of ethylene oxide. The products with M(n) up to 10(5) and having up to approximately 40 arms were obtained. The number of arms (f) was determined by direct measurements of M(n) of the first-generation stars (M(n) of arms A(x) is known), compared with f calculated from the branching index g, determined from R(g) measured with size-exclusion chromatography (SEC) triple detection with TriSEC software. The progress of the star formation was monitored by 1H NMR and SEC. These novel water-soluble stars, having a large number of hydroxyl groups, both at the ends of PEO arms as well as within the PGL arms, can be functionalized and further used for attaching compounds of interest. This approach opens, therefore, a new way of "multiPEGylation".  相似文献   

16.
Adenylate kinase from Escherichia coli (AKeco), consisting of a 23.6-kDa polypeptide chain folded into domains CORE, AMPbd, and LID catalyzes the reaction AMP + ATP <--> 2ADP. The domains AMPbd and LID execute large-amplitude movements during catalysis. Backbone dynamics of ligand-free and AP(5)A-inhibitor-bound AKeco is studied with slowly relaxing local structure (SRLS) (15)N relaxation, an approach particularly suited when the global (tau(m)) and the local (tau) motions are likely to be coupled. For AKeco tau(m) = 15.1 ns, whereas for AKeco*AP(5)A tau(m) = 11.6 ns. The CORE domain of AKeco features an average squared order parameter, , of 0.84 and correlation times tau(f) = 5-130 ps. Most of the AKeco*AP(5)A backbone features = 0.90 and tau(f) = 33-193 ps. These data are indicative of relative rigidity. Domains AMPbd and LID of AKeco, and loops beta(1)/alpha(1), alpha(2)/alpha(3), alpha(4)/beta(3), alpha(5)/beta(4), and beta(8)/alpha(7) of AKeco*AP(5)A, feature a novel type of protein flexibility consisting of nanosecond peptide plane reorientation about the C(i-1)(alpha)-C(i)(alpha) axis, with correlation time tau(perpendicular) = 5.6-11.3 ns. The other microdynamic parameters underlying this dynamic model include S(2) = 0.13-0.5, tau(parallel) on the ps time scale, and a diffusion tilt beta(MD) ranging from 12 to 21 degrees. For the ligand-free enzyme the tau(perpendicular) mode was shown to represent segmental domain motion, accompanied by conformational exchange contributions R(ex) < or = 4.4 s(-1). Loop alpha(4)/beta(3) and alpha(5)/beta(4) dynamics in AKeco*AP(5)A is related to the "energetic counter-balancing of substrate binding" effect apparently driving kinase catalysis. The other flexible AKeco*AP(5)A loops may relate to domain motion toward product release.  相似文献   

17.
The secondary structure of the purified 70-kDa protein Na+/Ca2+ exchanger, functionally reconstituted into asolectin lipid vesicles, was examined by Fourier transform infrared attenuated total reflection spectroscopy. Fourier transform infrared attenuated total reflection spectroscopy provided evidence that the protein is composed of 44% alpha-helices, 25% beta-sheets, 16% beta-turns, and 15% random structures, notably the proportion of alpha-helices is greater than that corresponding to the transmembrane domains predicted by exchanger hydropathy profile. Polarized infrared spectroscopy showed that the orientation of helices is almost perpendicular to the membrane. Tertiary structure modifications, induced by addition of Ca2+, were evaluated by deuterium/hydrogen exchange kinetic measurements for the reconstituted exchanger. This approach was previously proven as a useful tool for detection of tertiary structure modifications induced by an interaction between a protein and its specific ligand. Deuterium/hydrogen exchange kinetic measurements indicated that, in the absence of Ca2+, a large fraction of the protein (40%) is inaccessible to solvent. Addition of Ca2+ increased to 55% the inaccessibility to solvent, representing a major conformational change characterized by the shielding of at least 93 amino acids.  相似文献   

18.
A series of mononuclear copper(II) complexes having a 1:1 molar ratio of copper and the planar heterocyclic base like 1,10-phenanthroline (phen), dipyrido[3,2-d:2',3'-f]quinoxaline (dpq) and dipyrido[3,2-a:2',3'-c]phenazine (dppz) are prepared from a reaction of copper(II) nitrate.trihydrate and the base (L) in ethanol or aqueous ethanol at different temperatures. The complexes [Cu(dpq)(NO(3))(2)] (2), [Cu(dpq)(NO(3))(H(2)O)(2)](NO(3)) (3), [Cu(dpq)(NO(3))(2)(H(2)O)(2)].2H(2)O (4.2H(2)O) and [Cu(dppz)(NO(3))(2)(H(2)O)].H(2)O (5.H(2)O) have been characterized by X-ray crystallography. The crystal structures show the presence of the heterocyclic base in the basal plane. The coordination geometries of the copper(II) centers are axially elongated square-pyramidal (4+1) in 2, 3 and 5, and octahedral (4+2) in 4. The nitrate anion in the coordination sphere displays unidentate and bidentate chelating bonding modes. The axial ligand is either H(2)O or NO(3) in these structures giving a Cu-L(ax) distance of approximately 2.4 A. The one-electron paramagnetic complexes (mu approximately 1.8 mu(B)) exhibit axial EPR spectra in DMF glass at 77 K giving g(parallel)>g( perpendicular ) with an A(parallel) value of approximately 170G indicating a [d(x)2(-y)2](1) ground state. The complexes are redox active and display a quasireversible cyclic voltammetric response for the Cu(II)/Cu(I) couple near 0.0 V vs. SCE giving an order of the E(1/2) values as 5(dppz)>2-4 (dpq)>[Cu(phen)(2)(H(2)O)](2+)>1 (phen). The complexes bind to calf thymus DNA giving an order 5 (dppz)>2 (dpq)>[Cu(phen)(2)(H(2)O)](2+)>1 (phen). An effect of the extended planar ring in dpq and dppz is observed in the DNA binding. The complexes show nuclease activity with pUC19 supercoiled DNA in DMF/Tris-HCl buffer containing NaCl in presence of mercaptopropanoic acid as a reducing agent. The extent of cleavage follows the order: [Cu(phen)(2)(H(2)O)](ClO(4))(2)>5>2 approximately 3 approximately 4>1. The bis-phen complex is a better cleaver of SC DNA than 1-5 having mono-heterocyclic base. Mechanistic investigations using distamycin reveal minor groove biding for the phen, dpq complexes, and a major groove binding for the dppz complex 5. The cleavage reactions are found to be inhibited in the presence of hydroxyl radical scavenger DMSO and the reactions are proposed to proceed via sugar hydrogen abstraction pathway. The ancillary ligand is found to have less effect in DNA binding but are of importance in DNA cleavage reactions.  相似文献   

19.
A mathematical problem relating to membrane cylinders is stated and solved; its implications are illustrated and discussed. The problem concerns the volume distribution, in cylindrical coordinates, of the electric potential inside and outside a membrane cylinder of finite length (with sealed ends), during passive decay of an initially nonuniform membrane potential. The time constants for equalization with respect to the angle, theta, are shown to be typically about ten thousand times smaller than the time constant, tau(m) = R(m)C(m), for uniform passive membrane potential decay. The time constants for equalization with respect to length are shown to agree with those from one-dimensional cable theory; typically, they are smaller than tau(m) by a factor between 2 and 10. The relation of the membrane current density, I(m)(theta, x, t), to the values (at the outer membrane surface) of the extracellular potential phi(e)(r, theta, x, t) and of partial differential(2)phi(e)/ partial differentialx(2), is examined and it is shown that these quantities are not proportional to each other, in general; however, under certain specified conditions, all three of these quantities are proportional with each other and with phi(i)(r, theta, x, t) and partial differential(2)phi(i)/ partial differentialx(2) (at the inner membrane surface). The relation of these results to those of one-dimensional cable theory is discussed.  相似文献   

20.
The outer membrane (OM) ofErwinia amylovora was separated from the cytoplasmic membrane either by isopycnic sucrose density gradient centrifugation or by treating the envelope preparation with sodium lauroylsarcosine. Outer membranes, prepared by using either method, were similar in the content of the major proteins as revealed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The wild-type strains ofE. amylovora (E9, E8, EA178, EA198, EA225, EA273) had—in common in the OM—two major protein bands of apparent molecular weights of 15,800 (protein z) and 38,000 (protein y). The OM protein profiles of the virulent wild-type (E9) and the avirulent mutant (E8) were identical but differed from otherE. amylovora strains by the presence of an additional major protein band of approximately 40,000 (protein x). In strain E9, protein x appeared to be associated with the peptidoglycan, whereas proteins y and z were apparently not peptidoglycan-bound.  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