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 88 毫秒
1.
The rates of deuterium exchange reactions of malondialdehyde (MDA) and deuterated malondialdehyde (MDAd) have been studied as a function of acidity and the content of dimethyl sulfoxide (DMSO) in binary mixtures with D2O . MDA incorporates deuterium from D2O solutions in a first-order reaction with a rate constant (kobs) that depends on the acid concentration. From this dependence, a catalytic constant, kcat, can be derived (kcatMDA = 2.25 × 105M?s?1). Similar kinetic behavior was found for MDAd in H2O solutions, and in this case, kcatMDA = 1.56 × 105M?1s?1. Results from reactions of MDA and MDAd in identical H2OD2O mixtures show that primary and secondary isotope effects are small (kH/kD = 1.13) and that solvent isotope effects cause most of the differences found between reactions in D2O and H2O. Reactions in binary DMSOd6D2O mixtures show a six-fold rate increase as the proportion of DMSOd6 increases from 50% to 90%. These results also illustrate the relatively high reactivity of MDA at pH values well above its pKa and the importance of medium composition on its reaction rate.  相似文献   

2.
Protein aggregation leading to formation of amyloid fibrils is a symptom of several diseases like Alzheimer’s, type 2 diabetes and so on. Elucidating the poorly understood mechanism of such phenomena entails the difficult task of characterizing the species involved at each of the multiple steps in the aggregation pathway. It was previously shown by us that spontaneous aggregation of hen-eggwhite lysozyme (HEWL) at room temperature in pH 12.2 is a good model to study aggregation. Here in this paper we investigate the growth kinetics, structure, function and dynamics of multiple intermediate species populating the aggregation pathway of HEWL at pH 12.2. The different intermediates were isolated by varying the HEWL monomer concentration in the 300 nM—0.12 mM range. The intermediates were characterized using techniques like steady-state and nanosecond time-resolved fluorescence, atomic force microscopy and dynamic light scattering. Growth kinetics of non-fibrillar HEWL aggregates were fitted to the von Bertalanffy equation to yield a HEWL concentration independent rate constant (k = (6.6±0.6)×10−5 s−1). Our results reveal stepwise changes in size, molecular packing and enzymatic activity among growing HEWL aggregates consistent with an isodesmic aggregation model. Formation of disulphide bonds that crosslink the monomers in the aggregate appear as a unique feature of this aggregation. AFM images of multiple amyloid fibrils emanating radially from amorphous aggregates directly confirmed that on-pathway fibril formation was feasible under isodesmic polymerization. The isolated HEWL aggregates are revealed as polycationic protein nanoparticles that are robust at neutral pH with ability to take up non-polar molecules like ANS.  相似文献   

3.
The relationship between the sedimentation coefficient s0 and its concentration coefficient ks obtained in experiments on velocity sedimentation for polysaccharides is discussed. The values of s0, ks and an independently determined molecular weight reported by different authors for different polysaccharides are considered. It was established that the scaling relation. ks∼ s0 v unambiguously relates to the scaling relation s0∼ Mb. The values of the sedimentation parameter β s introduced on the basis of Svedberg's equation for s0 and on the basis of the expression ks = B 〈h23/2M–1 are discussed and the generalized Wales-van Holde-Rowe equation MKS = (NA/β S)3/2[s]3/2 kS 1/2 is used for evaluation of the molecular weights of polysaccharides. The adequacy of this evaluation is illustrated by taking as an example the determination of the unit length weight of an extra-rigid polysaccharide chain and of the equilibrium rigidity of rigid-chain, semi-rigid-chain and flexible-chain polysaccharides. The pair of experimental values s0 and kS obtained in a single series of experiments give the same information as may be obtained from the other pairs of hydrodynamic values such as [η] and s0 or [η] and D0, where [η] is the intrinsic viscosity and D0 is the translational diffusion coefficient. Accepted: 11 December 1996  相似文献   

4.
A laboratory-made sample of the polysaccharide xylinan (acetan) has been further characterized with respect to (i) purity, (ii) molar mass and polydispersity, and (iii) gross conformation by a combination of hydrodynamic measurements (sedimentation velocity and equilibrium analytical ultracentrifugation, viscometry, and dynamic light scattering) in aqueous NaCl (I = 0.10 mol·L−1). Sedimentation velocity diagrams recorded using Schlieren optics revealed highly pure material sedimenting as a single boundary [so20.w = 9.5 ± 0.7) S; ks = (273 ± 112) mL/g]. The hypersharp nature of these boundaries is symptomatic of a polydisperse and highly nonideal (in the thermodynamic sense) system. Low speed sedimentation equilibrium in the analytical ultracentrifuge using Rayleigh interference optics and two different types of extrapolation procedure (involving point and whole-cell molar masses) gave a weight average molar mass Mw of (2.5 ± 0.5) × 10−6 g·mol−1 and also a second virial coefficient, B = (2.8 ± 0.7) × 10−4 mL·mol·g−2, both values in good agreement with those from light scattering-based procedures (Part II of this series). A dynamic Zimm plot from dynamic light scattering measurements gave a z-average translational diffusion coefficient Do20.w = (3.02 ± 0.05) × 10−8 cm2·s−1 and the concentration-dependence parameter kD = (370 ± 15) mL/g. Combination of so20.w with Do20.w via the Svedberg equation gave another estimate for Mw of ≅ 2.4 × 106 g/mol, again in good agreement. Both the Wales-van Holde ratio (ks/[η]) ≅ 0.4 (with [η] = (760 ± 77) mL/g) and the ρ-parameter (ratio of the radius of gyration from static light scattering to the hydrodynamic radius from dynamic light scattering) as ρ > 2.0 all indicate an extended conformation for the macromolecules in solution. These findings, plus Rinde-type simulations of the sedimentation equilibrium data are all consistent with the interpretation in terms of a unimodal wormlike coil model performed earlier. © 1996 John Wiley & Sons, Inc.  相似文献   

5.
The kinetics of Ca2+-dependent conformational changes of human cardiac troponin (cTn) were studied on isolated cTn and within the sarcomeric environment of myofibrils. Human cTnC was selectively labeled on cysteine 84 with N-((2-(iodoacetoxy)ethyl)-N-methyl)amino-7-nitrobenz-2-oxa-1,3-diazole and reconstituted with cTnI and cTnT to the cTn complex, which was incorporated into guinea pig cardiac myofibrils. These exchanged myofibrils, or the isolated cTn, were rapidly mixed in a stopped-flow apparatus with different [Ca2+] or the Ca2+-buffer 1,2-Bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid to determine the kinetics of the switch-on or switch-off, respectively, of cTn. Activation of myofibrils with high [Ca2+] (pCa 4.6) induced a biphasic fluorescence increase with rate constants of >2000 s−1 and ∼330 s−1, respectively. At low [Ca2+] (pCa 6.6), the slower rate was reduced to ∼25 s−1, but was still ∼50-fold higher than the rate constant of Ca2+-induced myofibrillar force development measured in a mechanical setup. Decreasing [Ca2+] from pCa 5.0-7.9 induced a fluorescence decay with a rate constant of 39 s−1, which was approximately fivefold faster than force relaxation. Modeling the data indicates two sequentially coupled conformational changes of cTnC in myofibrils: 1), rapid Ca2+-binding (kB ≈ 120 μM−1 s−1) and dissociation (kD ≈ 550 s−1); and 2), slower switch-on (kon = 390s−1) and switch-off (koff = 36s−1) kinetics. At high [Ca2+], ∼90% of cTnC is switched on. Both switch-on and switch-off kinetics of incorporated cTn were around fourfold faster than those of isolated cTn. In conclusion, the switch kinetics of cTn are sensitively changed by its structural integration in the sarcomere and directly rate-limit neither cardiac myofibrillar contraction nor relaxation.  相似文献   

6.
An immobilization procedure for protein on surface plasmon resonance sensor (SPR) chips is described. The target protein, cyclophilin D, is thereby genetically linked to a mutant of the human DNA repair protein O6-alkylguanine-DNA-alkyltransferase (hAGT). The procedure includes the immobilization of an alkylguanine derivative on the surface by amine coupling and contact of the surface with a solution of the fusion protein (TCypD-hAGT). TCypD-hAGT could be immobilized using buffer solutions of purified protein or cell extracts. High densities of covalently linked proteins were achieved by either procedure. Binding experiments performed with the ligand cyclosporin A indicate relative binding activities close to 100%. The KD value (12 nM) and the kinetic rate constants kon (3 × 105 M−1s−1) and koff (4 × 10−3s−1) are given and compared to values determined for cyclophilin D linked to the surface by amide coupling chemistry. The KD value is in excellent agreement with the KD value determined in solution by fluorescence titration.  相似文献   

7.
The oligomerization of β-lactoglobulin (βLg) has been studied extensively, but with somewhat contradictory results. Using analytical ultracentrifugation in both sedimentation equilibrium and sedimentation velocity modes, we studied the oligomerization of βLg variants A and B over a pH range of 2.5–7.5 in 100 mM NaCl at 25°C. For the first time, to our knowledge, we were able to estimate rate constants (koff) for βLg dimer dissociation. At pH 2.5 koff is low (0.008 and 0.009 s−1), but at higher pH (6.5 and 7.5) koff is considerably greater (>0.1 s−1). We analyzed the sedimentation velocity data using the van Holde-Weischet method, and the results were consistent with a monomer-dimer reversible self-association at pH 2.5, 3.5, 6.5, and 7.5. Dimer dissociation constants KD2-1 fell close to or within the protein concentration range of ∼5 to ∼45 μM, and at ∼45 μM the dimer predominated. No species larger than the dimer could be detected. The KD2-1 increased as |pH-pI| increased, indicating that the hydrophobic effect is the major factor stabilizing the dimer, and suggesting that, especially at low pH, electrostatic repulsion destabilizes the dimer. Therefore, through Poisson-Boltzmann calculations, we determined the electrostatic dimerization energy and the ionic charge distribution as a function of ionic strength at pH above (pH 7.5) and below (pH 2.5) the isoelectric point (pI∼5.3). We propose a mechanism for dimer stabilization whereby the added ionic species screen and neutralize charges in the vicinity of the dimer interface. The electrostatic forces of the ion cloud surrounding βLg play a key role in the thermodynamics and kinetics of dimer association/dissociation.  相似文献   

8.
Reaction of nickel (II) perchlorate with the ligand N,N′-bis-(3,5-dipiperidin-1-yl-[2,4,6]triazin-1-yl)-pyridin-2-ylmethyl-ethane-1,2-diamine yields an octahedral Ni(II) high-spin complex 1 ([C40H56N14Ni(H2O)(CH3OH)](ClO4)2(CH3OH)2) with moderate zero-field splitting (ZFS) axial distortion parameter D/kB = 5.37 K. The ligand contributes a N4 donor set; the remaining two coordinating positions are occupied by coordinating solvents molecules. Exchange of the coordinating solvents molecules in complex 1 to thiocyanate moieties leads to formation of complex 2 ([C40H56N14Ni(NCS)2](CHCl)3) with an extended parameter D/kB = 8.80 K. The analysis of the structural and magnetic properties of complexes 1 and 2 led to the design of dinuclear complex 3 ([C40H56N14NiN3]2(ClO4)2(CH3OH)2), where two azido groups were utilized as bridging ligands. The double azido bridges in complex 3 cross each other to form a rarely observed non-coplanar (N3)2 structure. The magnetic behavior of complex 3 reveals ferromagnetic coupling interactions characterized by J/kB = 23.25 K, D1/kB = 7.90 K, D2/kB = 0.54 K.  相似文献   

9.
Static and dynamic light-scattering measurements are reported on zinc-insulin at room temperature (21 ± l°C) and pH = 6.88 in 0.1M NaCl aqueous solution. The experiments were performed at very low concentration, in the range 0.12 × 10?4 to 0.90 × 10?4 g cm?3. Within experimental error, we find no evidence for a critical micellar concentration in this system. The aggregation phenomenon starts immediately after preparation of the solutions, and takes several days to come to stable equilibrium. The concentration dependence of the diffusion coefficients, D z, = Do (1 — kDC), is negative, and kD was observed to decrease as a function of time, while the aggregate size was found to increase. The equivalent concentration coefficient, ?2BM W, obtained from static light scattering, showed a similar behavior, and, within experimental error, was found to be numerically equal to kD. From the relation found between the diffusion coefficient at infinite dilution and the molecular weight of the aggregates, log D0 = ?0.240 log M w ? 5.077, we deduce that the insulin aggregates are compact structures with a characteristic radius of 0.71 Å/(dalton)1/3, surrounded by a hydration layer of a thickness of 8.0 Å. The equilibrium aggregation number is approximately 10.  相似文献   

10.
Specific-ion effects are ubiquitous in nature; however, their underlying mechanisms remain elusive. Although Hofmeister-ion effects on proteins are observed at higher (>0.3M) salt concentrations, in dilute (<0.1M) salt solutions nonspecific electrostatic screening is considered to be dominant. Here, using effective charge (Q*) measurements of hen-egg white lysozyme (HEWL) as a direct and differential measure of ion-association, we experimentally show that anions selectively and preferentially accumulate at the protein surface even at low (<100 mM) salt concentrations. At a given ion normality (50 mN), the HEWL Q* was dependent on anion, but not cation (Li+, Na+, K+, Rb+, Cs+, GdnH+, and Ca2+), identity. The Q* decreased in the order F > Cl > Br > NO3 ∼ I > SCN > ClO4 ≫ SO42−, demonstrating progressively greater binding of the monovalent anions to HEWL and also show that the SO42− anion, despite being strongly hydrated, interacts directly with the HEWL surface. Under our experimental conditions, we observe a remarkable asymmetry between anions and cations in their interactions with the HEWL surface.  相似文献   

11.
The interactions between cytochrome P450 2B4 (d-2B4), NADPH:cytochrome P450 reductase and cytochrome b5 have been investigated in the monomeric reconstituted P450 2B4-containing monooxygenase system in the presence of a substrate (7-pentoxyresorufin) and an electron donor, NADPH. Each partner was immobilized via its amino groups on the carboxymethyldextran biochip surface of the optical biosensor IAsys+. Such mode immobilization was not accompanied by any loss of activities of the immobilized proteins. The formation of binary d-Fp/d-2B4 complexes was registered. The association/dissociation rate constants (kon/koff) were (0.013 ± 0.005) × 106 M?1 s?1/0.05 ± 0.02 s?1, and dissociation constant (KD) was (0.26 ± 0.13) × 10?6 M. Comparison of kon, koff and KD values for d-Fp/d-2B4 complexes formed under hydroxylation (O-dealkylation) with corresponding constants obtained for the oxidized proteins of (0.10 ± 0.03) × 106 M?1 s?1/(0.14 ± 0.06) s?1, and (0.71 ± 0.37) × 10?6 M, respectively shows that the decrease in kon and an insignificant decrease in KD are associated with the increase of complex lifetime during transition from the oxidized to hydroxylation conditions. Complex formation between d-Fp and d-b5 was not registered in both hydroxylation conditions and in the case of oxidized forms of these proteins. In both cases formation of the ternary d-Fp/d-2B4/d-b5 complexes occurred.  相似文献   

12.
We have investigated the potential of new methods of analysis of sedimentation velocity (SV) analytical ultracentrifugation (AUC) for the characterization of detergent-solubilized membrane proteins. We analyze the membrane proteins Ca++-ATPase and ExbB solubilized with DDM (dodecyl-β-d-maltoside). SV is extremely well suited for characterizing sample heterogeneity. DDM micelles (s 20w?=?3.1 S) and complexes (Ca++-ATPase: s 20w?=?7.3 S; ExbB: s 20w?=?4 S) are easily distinguished. Using different detergent and protein concentrations, SV does not detect any evidence of self-association for the two proteins. An estimate of bound detergent of 0.9 g/g for Ca++-ATPase and 1.5 g/g for ExbB is obtained from the combined analysis of SV profiles obtained using absorbance and interference optics. Combining s 20w with values of the hydrodynamic radius, R s?=?5.5 nm for Ca++-ATPase or R s?=?3.4 nm for ExbB, allows the determination of buoyant molar masses, M b. In view of their M b and composition, Ca++-ATPase and ExbB are monomers in our experimental conditions. We conclude that one of the main advantages of SV versus other techniques is the possibility to ascertain the homogeneity of the samples and to focus on a given complex even in the presence of other impurities or aggregates. The relative rapidity of SV measurements also allows experiments on unstable samples.  相似文献   

13.
Abstract— Proteins of the brain extracts of 85 individual pigeons (Columba livia) were mapped by two-dimensional gel electrophoresis. The method is a modification of O'Farrell 'S technique and separates proteins first by charge and then by molecular weight. There were three proteins, A, B and D which had each a variant form. The positions of these six proteins on the gel corresponded to the following pH values and molecular weight values: protein A1, 6.4/43,000; A2, 6.6/43,000; B1, 5.7/41,000; B2, 5.8/40,000; D1, 6.2/22,000; D2, 6.2/21,000. The variants are genetically determined, since protein A, B and D each occurred in three phenotypes (A1, A1A2 and A2; B1, B1B2 and B2; D1, D1D2 and D2) corresponding to the three possible genotypes. From the observed frequencies of the phenotypes the following allele frequencies were calculated: allele A1, 72%; A2, 28%; B1, 15%; B2, 85%; D1, 74%; D2, 26%. A fourth protein named C occurred in four different forms (C1, 7.2/37,000; C2, 7.2/36,000; C3, 7.1/37,000; C4, 7.1/36,000) and six phenotypes (C1, C1C2, C2, C1C3, C2C3 and C4C3). This polymorphism is also interpreted as being genetically determined. The four alleles coding for the four protein C forms had the following frequencies: allele C1, 62%; C2, 27%; C3, 10.5%; C4, 0.5%.  相似文献   

14.
NMR-monitored chemical shift titrations for the study of weak protein?Cligand interactions represent a rich source of information regarding thermodynamic parameters such as dissociation constants (K D ) in the micro- to millimolar range, populations for the free and ligand-bound states, and the kinetics of interconversion between states, which are typically within the fast exchange regime on the NMR timescale. We recently developed two chemical shift titration methods wherein co-variation of the total protein and ligand concentrations gives increased precision for the K D value of a 1:1 protein?Cligand interaction (Markin and Spyracopoulos in J Biomol NMR 53: 125?C138, 2012). In this study, we demonstrate that classical line shape analysis applied to a single set of 1H?C15N 2D HSQC NMR spectra acquired using precise protein?Cligand chemical shift titration methods we developed, produces accurate and precise kinetic parameters such as the off-rate (k off ). For experimentally determined kinetics in the fast exchange regime on the NMR timescale, k off ?~?3,000?s?1 in this work, the accuracy of classical line shape analysis was determined to be better than 5?% by conducting quantum mechanical NMR simulations of the chemical shift titration methods with the magnetic resonance toolkit GAMMA. Using Monte Carlo simulations, the experimental precision for k off from line shape analysis of NMR spectra was determined to be 13?%, in agreement with the theoretical precision of 12?% from line shape analysis of the GAMMA simulations in the presence of noise and protein concentration errors. In addition, GAMMA simulations were employed to demonstrate that line shape analysis has the potential to provide reasonably accurate and precise k off values over a wide range, from 100 to 15,000?s?1. The validity of line shape analysis for k off values approaching intermediate exchange (~100?s?1), may be facilitated by more accurate K D measurements from NMR-monitored chemical shift titrations, for which the dependence of K D on the chemical shift difference (????) between free and bound states is extrapolated to ?????=?0. The demonstrated accuracy and precision for k off will be valuable for the interpretation of biological kinetics in weakly interacting protein?Cprotein networks, where a small change in the magnitude of the underlying kinetics of a given pathway may lead to large changes in the associated downstream signaling cascade.  相似文献   

15.
This investigation examined the feasibility of manipulating the rotor speed in sedimentation velocity experiments to spontaneously generate an approximate steady-state condition where the extent of diffusional spreading is matched exactly by the boundary sharpening arising from negative sc dependence. Simulated sedimentation velocity distributions based on the sedimentation characteristics for a purified mucin preparation were used to illustrate a simple procedure for determining the diffusion coefficient from such steady-state distributions in situations where the concentration dependence of the sedimentation coefficient, s = s0/(1 + Kc), was quantified in terms of the limiting sedimentation coefficient as c → 0 (s0) and the concentration coefficient (K). Those simulations established that spontaneous generation of the approximate steady state could well be a feature of sedimentation velocity distributions for many unstructured polymer systems because the requirement that Kcoω2s0/D be between 46 and 183 cm−2 is not unduly restrictive. Although spontaneous generation of the approximate steady state is also a theoretical prediction for structured macromolecular solutes exhibiting linear concentration dependence of the sedimentation coefficient, s = s0(1 − kc), the required value of k is far too large for any practical advantage to be taken of this approach with globular proteins.  相似文献   

16.
The hemicellulose xylan constitutes a major portion of plant biomass, a renewable feedstock available for conversion to biofuels and other bioproducts. β-xylosidase operates in the deconstruction of the polysaccharide to fermentable sugars. Glycoside hydrolase family 43 is recognized as a source of highly active β-xylosidases, some of which could have practical applications. The biochemical details of four GH43 β-xylosidases (those from Alkaliphilus metalliredigens QYMF, Bacillus pumilus, Bacillus subtilis subsp. subtilis str. 168, and Lactobacillus brevis ATCC 367) are examined here. Sedimentation equilibrium experiments indicate that the quaternary states of three of the enzymes are mixtures of monomers and homodimers (B. pumilus) or mixtures of homodimers and homotetramers (B. subtilis and L. brevis). k cat and k cat/K m values of the four enzymes are higher for xylobiose than for xylotriose, suggesting that the enzyme active sites comprise two subsites, as has been demonstrated by the X-ray structures of other GH43 β-xylosidases. The K i values for d-glucose (83.3–357 mM) and d-xylose (15.6–70.0 mM) of the four enzymes are moderately high. The four enzymes display good temperature (K t 0.5?~?45 °C) and pH stabilities (>4.6 to <10.3). At pH 6.0 and 25 °C, the enzyme from L. brevis ATCC 367 displays the highest reported k cat and k cat/K m on natural substrates xylobiose (407 s?1, 138 s?1?mM?1), xylotriose (235 s?1, 80.8 s?1?mM?1), and xylotetraose (146 s?1, 32.6 s?1?mM?1).  相似文献   

17.
Transferrin receptor 1 (RD) binds iron-loaded transferrin and allows its internalization in the cytoplasm. Human serum transferrin also forms complexes with metals other than iron, including uranium in the uranyl form (UO2 2+). Can the uranyl-saturated transferrin (TUr2) follow the receptor-mediated iron-acquisition pathway? In cell-free assays, TUr2 interacts with RD in two different steps. The first is fast, direct rate constant, k 1 = (5.2 ± 0.8) × 106 M?1 s?1; reverse rate constant, k ?1 = 95 ± 5 s?1; and dissociation constant K 1 = 18 ± 6 μM. The second occurs in the 100-s range and leads to an increase in the stability of the protein–protein adduct, with an average overall dissociation constant K d = 6 ± 2 μM. This kinetic analysis implies in the proposed in vitro model possible but weak competition between TUr2 and the C-lobe of iron-loaded transferrin toward the interaction with R D.  相似文献   

18.
The xyn10B gene, encoding the endo-1,4-β-xylanase Xyn10B from Thermotoga thermarum, was cloned and expressed in Escherichia coli. The ORF of the xyn10B was 1,095 bp and encoded to mature peptide of 344 amino acids with a calculated MW of 40,531 Da. The recombinant xylanase was optimally active at 80 °C, pH 6.0 and retained approx. 60 % of its activity after 2 h at 75 °C. Apparent K m , k cat and k cat /K m values of the xylanase for beechwood xylan were 1.8 mg ml?1, 520 s?1 and 289 ml mg?1 s?1, respectively. The end products of the hydrolysis of beechwood xylan were mainly oligosaccharides but without xylose after 2 h hydrolysis.  相似文献   

19.
The Lys residues in the 75-residue Ca2+-binding protein calbindin D9k were reductively methylated with13C-enriched formaldehyde. The possible structural effects resulting from the chemical modification were critically investigated by comparing two-dimensional NMR spectra and the exchange rates of some of the amide protons of the native and the modified protein. Our results show that the protein retains its structure even though 10 Lys out of a total of 75 amino acid residues were modified. In the Ca2+- and apo-forms of the protein, the13C-methylated Lys residues can be detected with high sensitivity and resolution using two-dimensional (1H,13C)-heteronuclear multiple quantum coherence (HMQC) NMR spectroscopy. ThepKa values of the individual Lys residues in Ca2+-calbindin D9k and apo-calbindin D9k were obtained by combiningpH titration experiments and (1H,13C)-HMQC NMR spectroscopy. Each Lys residue in the Ca2+- and apo-forms of calbindin D9k has a uniquepKa value. The LyspKa values in the calcium protein range from 9.3 to 10.9, while those in the apo-protein vary between 9.7 and 10.7. Although apo-calbindin D9k has a very similar structure compared to Ca2+-calbindin D9k, the removal of two Ca2+ ions from the protein leads to an increase of thepKa values of the Lys residues.  相似文献   

20.
A Model for the Action of Vinblastine in Vivo   总被引:1,自引:0,他引:1       下载免费PDF全文
A model for the action of vinblastine (VLB) on cells multiplying exponentially in vivo with a generation time, TG, has been derived. It is based on the assumption that cells attempting to pass through mitosis in the presence of VLB lose their proliferative capacity and that this lethal effect occurs only when the cells are exposed to a concentration of VLB which is above a critical value, Ck. The model leads to two predictions. First, that the percentage of cells surviving at any time after exposure to a dose, D, of VLB is 100% if D < Dk and decreases to 0% after a time, TG, following a dose D ≥ Dk·2T G/T1/2, where Dk represents the dose of VLB required to produce the concentration Ck, and T1/2 is the half-life of the VLB in vivo. Second, that the time, TG, at which the percentage of cells surviving an exposure to VLB, at doses greater than Dk·2U G/T1/2, decreases to zero should be equal to the generation time of the cells. Both of these predictions were confirmed experimentally which indicates that the model adequately explains the action of VLB in vivo.  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