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 32 毫秒
1.
Conformational states of ribulosebisphosphate carboxylase (Rubisco) from Rhodospirillum rubrum were examined by far-UV circular dichroism (CD), tryptophan fluorescence, and 1-anilino-naphthalenesulfonate (ANS) binding. At pH 2 and low ionic strength (I = 0.01), Rubisco adopts an unfolded, monomeric conformation (UA1 state) as judged by far-UV CD and tryptophan fluorescence. As with other acid-unfolded proteins [Goto, Y., Calciano, L. J., & Fink, A. L. (1990) Proc. Natl. Acad. Sci. U.S.A. 87, 573-577], an intermediate conformation (A1 state) is observed at pH 2 and high ionic strength. The A1 state has an alpha-helical content equivalent to 64% of that present in the native dimer (N2 state). However, fluorescence measurements indicate that the tertiary structure of the A1 state is largely disordered. A site-directed mutant, K168E, which exists as a stable monomer [Mural, R. J., Soper, T. S., Larimer, F. W., & Hartman, F. C. (1990) J. Biol. Chem. 265, 6501-6505] was used to characterize the "native" monomer (N1 state). The far-UV CD spectra of the N1 and N2 states are almost identical, indicating a similar secondary structure content. However, the tertiary structure of the N1 state is less ordered than that of the N2 state. Nevertheless, when appropriately complemented in vitro, K168E forms an active heterodimer. Upon neutralization of acid-denatured Rubisco or dilution of guanidine hydrochloride-denatured Rubisco, unstable folding intermediates (I1 state) are rapidly formed. At concentrations at or below the "critical aggregation concentration" (CAC), the I1 state reverts spontaneously but slowly to the native states with high yield (greater than 65%). The CAC is temperature-dependent. At concentrations above the CAC, the I1 and the A1 states undergo irreversible aggregation. The commitment to aggregation is rapid [ef. Goldberg, M. E., Rudolph, R., & Jaenicke, R. (1991) Biochemistry 30, 2790-2797] and proceeds until the concentration of folding intermediate(s) has fallen to the CAC. In the presence of a molar excess of chaperonin 60 oligomers, the I1 state forms a stable binary complex. No stable binary complex between chaperonin 60 and the N1 state could be detected. Formation of the chaperonin 60-I1 binary complex arrests the spontaneous folding process. The I1 state becomes resistant to interaction with chaperonin 60 with kinetics indistinguishable from those associated with the appearance of the native states. In vitro complementation analysis indicated that the product of the chaperonin-facilitated process is monomeric.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
Two previously unreported diploid numbers, 2N = 50 and 2N = 51, from five individual Lemur fulvus collaris are described. In both chromosome complements, the nombre fondamental is 64. The 2N = 50 complement is composed of 7 pairs of bi-armed chromosomes, 17 pairs of acrocentric chromosomes with small short arms, and acrocentric sex chromosomes. The 2N = 51 complement is identical with these exceptions: Only one member equivalent to bi-armed pair 6 of L.f. collaris (2N = 50) is present, and two extra acrocentrics are found in the 2N = 51 complement. G-banding analyses suggest that these chromosomes are a heteromorphic pair of the Robertsonian type. This conculsion is supported by evidence from studies of meiotic pairing relationships of the three chromosomes and the complements of hybrids resulting from interspecific matings. Comparison of the 2N = 50 and 2N = 51 complements with a published 2N = 48 complement suggests that these new karyotypes do not provide a lineal link between the 2N = 52 and 2N = 48 karyotypes (Rumpler, Y., and R. Albignac 1969 C.R. Soc. Biol., 163: 1989-1992).  相似文献   

3.
A low-molecular-mass serine protease inhibitor was purified from hepatocytes and liver of rats. It was found to be a single polypeptide of 56 amino acid residues corresponding to Mr = 6224, a value that is in agreement with the molecular mass determined by gel chromatography. The inhibitor formed a complex in a molar ratio of 1:1 with trypsin. Its complete amino acid sequence was identical with that of pancreatic secretory trypsin inhibitor II (PSTI-II) in pancreatic juice, but not with that of PSTI-I [Uda, K., Ogawa, M., Shibata, T., Murata, A., Mori, T., Kikuchi, N., Yoshida, N., Tsunasawa, S. & Sakiyama, F. (1988) Biol. Chem. Hoppe-Seyler 369, 55-61]. PSTIs have been reported to be primarily pancreatic secretory products, but in have been reported to be primarily pancreatic secretory products, but in patients immunoreactive PSTI was found in the plasma and urine during acute inflammatory disease and shown to be produced ectopically in cancer tissues. Here we report for the first time that PSTI-II is present in other normal tissues besides the pancreas.  相似文献   

4.
H M Chen  V S Markin  T Y Tsong 《Biochemistry》1992,31(49):12369-12375
Staphylococcal nuclease unfolds at acidic pHs and refolds at neutral pH. Previous kinetic analysis based on both the direct pH jump and the sequential pH jump, from a native condition (pH 7.0) to pHs beyond unfolding transition zones (pH 3.0 and pH 12), and vice versa, supports the mechanism, D3<-->D2<-->D1<-->N0, in which N0 is the native state and D's are the three substates of the denatured form [Chen, H.M., You, J.L., Markin, V.S., & Tsong, T.Y. (1990) J. Mol. Biol. 220, 771-778; Chen, H.M., Markin, V.S., & Tsong, T.Y. (1992) Biochemistry 31, 1483-1491]. Here we show that both the single- and the double-pH jump kinetics of folding and unfolding to the intermediate pHs (3.4-5.0, i.e., in the transition zone), in which both the native and the denatured states coexist, are not compatible with this simple sequential model. At 25 degrees C, log tau 1(-1) (for the D1<-->N0 step) and log tau 2(-1) (for the D2<-->D1 step) vs pH show a square root of-shaped dependence on the final pH, with minimal values (tau 1(-1) of 0.56 s-1 and tau 2(-1) of around pH 3.9. The third relaxation tau 3 (for the D3<-->D2 step, 35 s) was independent of pH in the range 3.4-8.5. The square root of-shaped dependence on pH of log tau 1(-1) and log tau 2(-1) cannot be reproduced by the above but can be accounted for if each of N0, D1, and D2 is composed of many microscopic states in rapid equilibrium.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Summary The influence of temperature (15–32°C) and the ratio of nitrogen to phosphorus (N/P) in the culture medium (0.5–80) on the growth kinetics and protein, chlorophyll, lipid and fatty acid content of the marine microalga Tetraselmis sp. have been studied. Below an N/P of 20, growth was determined by N limitation and above 20 by P limitation. Protein increased with a rise in N content at any test temperature. The chlorophyll content increased with temperature, with maximum values at 25°C. The lipid content decreased with increasing N/P ratio above 20°C. The polyunsaturated fatty acid content tends to be inversely proportional to the growth rate within the N/P range 20–80. The quotient of the n 3 and n 6 polyunsaturated-fatty-acid fractions, an indicator of the nutritive value of microalgae, was found to be within the range 2–3. These values were obtained either between 25 and 28°C independent of the N/P ratio used at 20°C for N/P ratios higher than 40.0. Offsprint requests to: Emilio Molina  相似文献   

6.
C Tesi  F Travers  T Barman 《Biochemistry》1988,27(13):4903-4908
The kinetics of the interaction of the fluorescent analogue 1,N6-ethenoadenosine 5'-triphosphate (epsilon-ATP) with myosin subfragment 1 (S1) were studied at 15 and -7.5 degrees C with 40% ethylene glycol as cryosolvent. Two techniques were used: fluorescence stopped flow and rapid flow-quench. When S1 is mixed with epsilon-ATP in a stopped-flow apparatus, biphasic fluorescence transients are obtained which are difficult to assign. Chemical sampling by the rapid-flow-quench method led to the chemical identity and the kinetics of interconversion of key intermediates, and by this method the optical signals were assigned and information about the cleavage and release of products was obtained. The data were interpreted by a shortened form of the Bagshaw-Trentham scheme for myosin adenosinetriphosphatase: M + ATP K1 in equilibrium M.ATP k2----M*.ATP k3 in equilibrium k3 M**.ADP.Pi k4----M + ADP + Pi The constants obtained were compared with those for ATP under identical conditions. In agreement with Rosenfeld and Taylor [Rosenfeld, S. S., & Taylor, E. W. (1984) J. Biol. Chem. 259, 11920-11929] we find that epsilon-ATP is bound tightly to S1 and that the chemical step is slower than with ATP. We show that the fast fluorescence transient is due to the tight binding of epsilon-ATP with K1 = 32 microM and k2 = 58 s-1 at 15 degrees C. With ATP these values are 8 microM and 16 s-1, respectively. There is a large difference in the delta H for k2: 50 kJ.mol-1 for epsilon-ATP and 119 kJ.mol-1 for ATP.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Grand canonical Monte Carlo (GCMC) simulations are reported for aqueous solutions containing excess univalent salt (activities a +/- = 1.76-12.3 mM) and one of the following species: an octacationic rod-like ligand, L8+; a B-DNA oligomer with N phosphate charges (8 < or = N < or = 100); or a complex resulting from the binding of L8+ at the center of an N-mer (24 < or = N < or = 250). Simplified models of these multiply charged species are used in the GCMC simulations to predict the fundamental coulombic contributions to the following experimentally relevant properties: 1) the axial distance over which ligand binding affects local counterion concentrations at the surface of the N-mer; 2) the dependence on N of GCMC preferential interaction coefficients, gamma 32MC identical to delta C3/delta C2l a +/-, T, where C3 and C2 are, respectively, the molar concentrations of salt and the multiply charged species (ligand, N-mer or complex); and 3) the dependence on N of SaKobs identical to d in Kobs/d in a +/- = delta (magnitude of ZJ + 2 gamma 32J), where Kobs is the equilibrium concentration quotient for the binding of L8+ to the center of an N-mer and delta denotes the stoichiometric combination of terms, each of which pertains to a reactant or product J having magnitude of ZJ charges. The participation of electrolyte ions in the ligand binding interaction is quantified by the magnitude of SaKobs, which reflects the net (stoichiometrically weighted) difference in the extent of thermodynamic binding of salt ions to the products and reactants. Results obtained here from GCMC simulations yield a picture of the salient molecular consequences of binding a cationic ligand, as well as thermodynamic predictions whose applicability can be tested experimentally. Formation of the central complex is predicted to cause a dramatic reduction in the surface counterion (e.g., Na+) concentration over a region including but extending well beyond the location of the ligand binding site. For binding a cationic ligand, SaKobs is predicted to be negative, indicating net electrolyte ion release in the binding process. At small enough N, -SaKobs is predicted to decrease strongly toward zero with decreasing N. At intermediate N, -SaKobs appears to exceed its limiting value as N-->infinity.  相似文献   

8.
The bacteriorhodopsin photocycle contains more than five spectrally distinct intermediates, and the complexity of their interconversions has precluded a rigorous solution of the kinetics. A representation of the photocycle of mutated D96N bacteriorhodopsin near neutral pH was given earlier (Váró, G., and J. K. Lanyi. 1991. Biochemistry. 30:5008-5015) as BRhv-->K<==>L<==>M1-->M2--> BR. Here we have reduced a set of time-resolved difference spectra for this simpler system to three base spectra, each assumed to consist of an unknown mixture of the pure K, L, and M difference spectra represented by a 3 x 3 matrix of concentration values between 0 and 1. After generating all allowed sets of spectra for K, L, and M (i.e., M1 + M2) at a 1:50 resolution of the matrix elements, invalid solutions were eliminated progressively in a search based on what is expected, empirically and from the theory of polyene excited states, for rhodopsin spectra. Significantly, the average matrix values changed little after the first and simplest of the search criteria that disallowed negative absorptions and more than one maximum for the M intermediate. We conclude from the statistics that during the search the solutions strongly converged into a narrow region of the multidimensional space of the concentration matrix. The data at three temperatures between 5 and 25 degrees C yielded a single set of spectra for K, L, and M; their fits are consistent with the earlier derived photocycle model for the D96N protein.  相似文献   

9.
Sequence-specific 15N and 1H assignments for the trp holorepressor from Escherichia coli are reported. The trp repressor consists of two identical 107-residue subunits which are highly helical in the crystal state [Schevitz, R., Otwinowski, Z., Joachimiak, A., Lawson, C. L. & Sigler, P. B. (1985) Nature 317, 782-786]. The high helical content and the relatively large size of the protein (Mr = 25,000) make it difficult to assign even the main-chain resonances by conventional homonuclear two-dimensional NMR methods. However, we have now assigned the main-chain resonances of 94% of the residues by using three-dimensional 15N/1H heteronuclear experiments on a sample of protein uniformly labelled with 15N. The additional resolution obtained by spreading out the signals into three dimensions proved indispensable in making these assignments. In particular, we have been able to resolve signals from residues in the N-terminal region of the A helix for the first time in solution. The observed NOE results confirm that the repressor is highly helical in solution, and contains no extended chain conformations.  相似文献   

10.
Single tryptophan-containing mutants of low adenylylation state Escherichia coli glutamine synthetase have been studied by frequency-domain fluorescence spectroscopy in the presence of various substrates and inhibitors. At pH 6.5, the Mn-bound wild-type enzyme (wild type has two tryptophans/subunit) and the mutant enzymes exhibit heterogeneous fluorescence decay kinetics; the individual tryptophans are adequately described by a triple exponential decay scheme. The recovered lifetime values are 5.9 ns, 2.6 ns, and 0.4 ns for Trp-57 and 5.8 ns, 2.3 ns, and 0.4 ns for Trp-158. These values are nearly identical to the previously reported results at pH 7.5 (Atkins, W.M., Stayton, P.S., & Villafranca, J.J., 1991, Biochemistry 30, 3406-3416). In addition, Trp-57 and Trp-158 both exhibit an ATP-induced increase in the relative fraction of the long lifetime component, whereas only Trp-57 is affected by this ligand at pH 7.5. The transition-state analogue L-methionine-(R,S)-sulfoximine (MSOX) causes a dramatic increase in the fractional intensity of the long lifetime component of Trp-158. This ligand has no effect on the W158S mutant protein and causes a small increase in the fractional intensity of the long lifetime component of the W158F mutant protein. Addition of glutamate to the ATP complex, which affords the gamma-glutamylphosphate-ADP complex, results in the presence of new lifetime components at 7, 3.2, and 0.5 ns for Trp-158, but has no effect on Trp-57. Similar results were obtained when ATP was added to the MSOX complex; Trp-57 exhibits heterogeneous fluorescence decay with lifetimes of 7, 3.5, and 0.8 ns. Decay kinetics of Trp-158 are best fit to a nearly homogeneous decay with a lifetime of 5.5 ns in the MSOX-ATP inactivated complex. These results provide a model for the sequence of structural and dynamic changes that take place at the Trp-57 loop and the central loop (Trp-158) during several intermediate stages of catalysis.  相似文献   

11.
Intravascular bubble formation and symptoms of decompression sickness have been reported during repetitive deep breath-hold diving. Therefore we examined the pattern of blood N2 kinetics during and after repetitive breath-hold diving. To study muscle N2 uptake and release, we measured brachial venous N2 partial pressure (PN2) in nine professional Korean breath-hold divers (ama) during a 3-h diving shift at approximately 4 m seawater depth and up to 4 h after diving. PN2 was determined with the manometric Van Slyke method. Diving time and depth were recorded using a backpack computer-assisted dive longer that allowed calculating the surface-to-depth time ratio to derive the effective depth. With the assumption that forearm muscle N2 kinetics follow the general Haldanian principles of compression and decompression, i.e., forearm muscle is a single compartment with a uniform tissue PN2 equal to venous PN2, PN2 data were fitted to monoexponential functions of time. In the early phase of the diving shift, PN2 rapidly increased to 640 Torr (half time = 6 min) and then slowly declined to baseline levels (half time = 36 min) after the work shift. Peak PN2 levels approximated the alveolar PN2 derived from the effective depth. We conclude that forearm muscle N2 kinetics are well described by a Haldanian single-compartment model. Decompression sickness is theoretically possible in the ama; it did not occur because the absolute PN2 remained low due to the shallow working depth of the ama we studied.  相似文献   

12.
The structure of the pneumococcal common antigen, C-polysaccharide, from a noncapsulated pneumococcal strain, CSR SCS2, was studied using 1H-NMR, 13C-NMR and 31P-NMR spectroscopy. The dependence of NMR chemical shifts on the variation in pD was also studied. It was established that the C-polysaccharide is composed of a backbone of tetrasaccharide-ribitol repeating units that are linked to each other by a phosphodiester linkage between position 5 of a D-ribitol residue and position 6 of a beta-D-glucopyranosyl residue. The polysaccharide is substituted with one residue of phosphocholine at position 6 of the 4-substituted 2-acetamido-2-deoxy-alpha-D-galactopyranosyl residue. Both galactosamine residues in the polysaccharide are N-acetylated. O)-P-Cho | 6 6)-beta-D-Glcp-(1-->3)-alpha-AATp-(1-->4)-alpha-D-GalpNAc-(1-->3)- bet a-D-GalpNAc-(1-->1)-D-ribitol-5-P-(O--> where AAT is 2-acetamido-4-amino-2,4,6-trideoxy-D-galactose and Cho is choline. This structure differs, concerning phosphocholine substituents and N-acetylation, from those reported previously for pneumococcal C-polysaccharide [Jennings, H.J., Lugowski, C. & Young, N.M. (1980) Biochemistry 19, 4712-4719; Fischer, W., Behr, T., Hartmann, R., Peter-Katalinic, J. & Egge, H. (1993) Eur. J. Biochem. 215, 851-857; Kulakowska, M., Brisson, J.-R., Griffith, D.W., Young, N.M. & Jennings, H.J. (1993) Can. J. Chem. 71, 644-648]. The structures of the C-polysaccharides present in three pneumococcal types were also examined. They contain one (in 18B) or two (in 32F and 32A) phosphocholine residues in the repeating unit. The degree of substitution was not determined. The backbone of all examined C-polysaccharides was identical and in all cases both galactosamine residues appeared to be N-acetylated.  相似文献   

13.
The molecular structures of three phosphorus-based peptide inhibitors of aspartyl proteinases complexed with penicillopepsin [1, Iva-L-Val-L-Val-StaPOEt [Iva = isovaleryl, StaP = the phosphinic acid analogue of statine [(S)-4-amino-(S)-3-hydroxy-6-methylheptanoic acid] (IvaVVStaPOEt)]; 2, Iva-L-Val-L-Val-L-LeuP-(O)Phe-OMe [LeuP = the phosphinic acid analogue of L-leucine; (O)Phe = L-3-phenyllactic acid; OMe = methyl ester] [Iva VVLP(O)FOMe]; and 3, Cbz-L-Ala-L-Ala-L-LeuP-(O)-Phe-OMe (Cbz = benzyloxycarbonyl) [CbzAALP(O)FOMe]] have been determined by X-ray crystallography and refined to crystallographic agreement factors, R ( = sigma parallel to F0 magnitude of - Fc parallel to/sigma magnitude of F0), of 0.132, 0.131, and 0.134, respectively. These inhibitors were designed to be structural mimics of the tetrahederal transition-state intermediate encountered during aspartic proteinase catalysis. They are potent inhibitors of penicillopepsin with Ki values of 1, 22 nM; 2, 2.8 nM; and 3, 1600 nM, respectively [Bartlett, P. A., Hanson, J. E., & Giannousis, P. P. (1990) J. Org. Chem. 55, 6268-6274]. All three of these phosphorus-based inhibitors bind virtually identically in the active site of penicillopepsin in a manner that closely approximates that expected for the transition state [James, M. N. G., Sielecki, A.R., Hayakawa, K., & Gelb, M. H. (1992) Biochemistry 31, 3872-3886]. The pro-S oxygen atom of the two phosphonate inhibitors and of the phosphinate group of the StaP inhibitor make very short contact distances (approximately 2.4 A) to the carboxyl oxygen atom, O delta 1, of Asp33 on penicillopepsin. We have interpreted this distance and the stereochemical environment of the carboxyl and phosphonate groups in terms of a hydrogen bond that most probably has a symmetric single-well potential energy function. The pro-R oxygen atom is the recipient of a hydrogen bond from the carboxyl group of Asp213. Thus, we are able to assign a neutral status to Asp213 and a partially negatively charged status to Asp33 with reasonable confidence. Similar very short hydrogen bonds involving the active site glutamic acid residues of thermolysin and carboxypeptidase A and the pro-R oxygen of bound phosphonate inhibitors have been reported [Holden, H. M., Tronrud, D. E., Monzingo, A. F., Weaver, L. H., & Matthews, B. W. (1987) Biochemistry 26, 8542-8553; Kim, H., & Lipscomb, W. N. (1991) Biochemistry 30, 8171-8180].(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
A direct continuous UV-Vis spectrophotometric assay has been developed for VanX, a D-alanyl-D-alanine aminodipeptidase necessary for vancomycin resistance. This method is based on the hydrolysis of the alternative substrate D-alanyl-alpha-(R)-phenylthio-glycine D-Ala-D-Gly(S-Ph)-OH (H-DAla-DPsg-OH, 5a). Spontaneous decomposition of the released phenylthioglycine generates thiophenol, which is quantified using Ellman's reagent. The dipeptide behaved as an excellent substrate of VanX, exhibiting Michaelis-Menten kinetics with a kcat of 76 +/- 5/s and a KM of 0.83 +/- 0.08 mm (kcat = 46 +/- 3/s and KM = 0.11 +/- 0.01 mm for D-Ala-D-Ala). Determination of the kinetic parameters of the previously reported mechanism-based inhibitor D-Ala-D-Gly(SPhip-CHF2)-OH (H-D-Ala-DPfg-OH, 5c) [Araoz, R., Anhalt, E., René, L., Badet-Denisot, M.-A., Courvalin, P. & Badet, B. (2000) Biochemistry 39, 15971-15979] using the substrate reported in the present study yielded values of Kirr of 22 +/- 1 microM and kinact of 9.3 +/- 0.4/min in good agreement with values previously obtained in our laboratory (Kirr = 30 +/- 1 mm; kinact = 7.3 +/- 0.3/min). In addition, inhibition by the competing substrate D-Ala-D-Ala resulted in determination of a Ki = 70 +/- 6 microM close to the previously reported KM value. These results demonstrate that the present assay is a convenient, rapid and sensitive tool in the search for VanX inhibitors.  相似文献   

15.
The binding of indirubin-5-sulphonate (E226), a potential anti-tumour agent and a potent inhibitor (IC(50) = 35 nm) of cyclin-dependent kinase 2 (CDK2) and glycogen phosphorylase (GP) has been studied by kinetic and crystallographic methods. Kinetic analysis revealed that E226 is a moderate inhibitor of GPb (K(i) = 13.8 +/- 0.2 micro m) and GPa (K(i) = 57.8 +/- 7.1 micro m) and acts synergistically with glucose. To explore the molecular basis of E226 binding we have determined the crystal structure of the GPb/E226 complex at 2.3 A resolution. Structure analysis shows clearly that E226 binds at the purine inhibitor site, where caffeine and flavopiridol also bind [Oikonomakos, N.G., Schnier, J.B., Zographos, S.E., Skamnaki, V.T., Tsitsanou, K.E. & Johnson, L.N. (2000) J. Biol. Chem.275, 34566-34573], by intercalating between the two aromatic rings of Phe285 and Tyr613. The mode of binding of E226 to GPb is similar, but not identical, to that of caffeine and flavopiridol. Comparative structural analyses of the GPb-E226, GPb-caffeine and GPb-flavopiridol complex structures reveal the structural basis of the differences in the potencies of the three inhibitors and indicate binding residues in the inhibitor site that can be exploited to obtain more potent inhibitors. Structural comparison of the GPb-E226 complex structure with the active pCDK2-cyclin A-E226 complex structure clearly shows the different binding modes of the ligand to GPb and CDK2; the more extensive interactions of E226 with the active site of CDK2 may explain its higher affinity towards the latter enzyme.  相似文献   

16.
G Váró  J K Lanyi 《Biochemistry》1991,30(20):5008-5015
The photocycles of wild-type bacteriorhodopsin and its D96N form were investigated with a gated multichannel analyzer. Reconstruction of the spectra of the photointermediates from the measured time-resolved difference spectra allowed evaluation of the kinetics; the data at pH 7 in the presence of 100 mM NaCl were best fitted by the scheme K in eqiulibrium L in equilibrium M1----M2 in equilibrium N in equilibrium O----BR plus N----BR [Váró, G., & Lanyi, J. K. (1990) Biochemistry 29, 2241-2250]. The proposed two M states and the M1----M2 reaction were necessitated by anomalies in the kinetics of the decay of K and L. Additional support was provided by a 4-nm blue-shift in the maximum of M in Triton X-100 solubilized bacteriorhodopsin during the photocycle; the kinetics of the shift were consistent with the time course of the proposed M1----M2 transition. In the D96N mutant, the M state is stabilized, and the resulting equilibrium mixture for the intermediates could be evaluated with greater precision. The concentration ratio of L to M at the equilibrium was estimated to be no higher than 0.01. This requires the ratio of forward/reverse rates for the M1 to M2 conversion to be at least 200, i.e., a virtually irreversible reaction. Consistent with an earlier report, the data at lower pH and in the absence of NaCl are different and suggest the existence of a second L species; we propose that it is in equilibrium with M2.  相似文献   

17.
Slips, trips, and falls are some of the most substantial and prevalent causes of occupational injuries and fatalities, and these events may contribute to low-back problems. We quantified lumbar kinematics (i.e., lumbar angles relative to pelvis) and kinetics during unexpected slip and trip perturbations, and during normal walking, among 12 participants (6F, 6 M). Individual anthropometry, lumbar muscle geometry, and lumbar angles, along with electromyography from 14 lumbar muscles were used as input to a 3D, dynamic, EMG-based model of the lumbar spine. Results indicated that, in comparison with values during normal walking, lumbar range of motion, lumbosacral (L5/S1) loads, and lumbar muscle activations were all significantly higher during the slip and trip events. Maximum L5/S1 compression forces exceeded 2700 N during slip and trip events, compared with ∼ 1100 N during normal walking. Mean values of L5/S1 anteroposterior (930 N), and lateral (800 N) shear forces were also substantially larger than the shear force during the normal walking (230 N). These observed levels of L5/S1 reaction forces, along with high levels of bilateral lumbar muscle activities, suggest the potential for overexertion injuries and tissue damage during unexpected slip and trip events, which could contribute to low back injuries. Outcomes of this study may facilitate the identification and control of specific mechanisms involved with low back disorders consequent to slips or trips.  相似文献   

18.
R K Harrison  R L Stein 《Biochemistry》1990,29(7):1684-1689
Cyclophilin, the cytosolic binding protein for the immunosuppressive drug cyclosporin A, has recently been shown to be identical with peptidyl prolyl cis-trans isomerase [Fischer, G., Wittmann-Liebold, B., Lang, K., Kiefhaber, T., & Schmid, F.X. (1989) Nature 337, 476; Takahashi, N., Hayano, T., & Suzuki, M. (1989) Nature 337, 473]. To provide a mechanistic framework for studies of the interaction of cyclophilin with cyclosporin, we investigated the mechanism of the PPI-catalyzed cis to trans isomerization of Suc-Ala-Xaa-cis-Pro-Phe-pNA (Xaa = Ala, Gly). Our mechanistic studies of peptidyl prolyl cis-trans isomerase include the determination of steady-state kinetic parameters, pH and temperature dependencies, and solvent and secondary deuterium isotope effects. The results of these experiments support a mechanism involving catalysis by distortion in which the enzyme uses free energy released from favorable, noncovalent interactions with the substrate to stabilize a transition state that is characterized by partial rotation about the C-N amide bond.  相似文献   

19.
Noncovalent aggregation is a side reaction in the process of reconstitution of oligomeric enzymes (e.g., lactic dehydrogenase) after preceding dissociation, denaturation, and deactivation. The aggregation product is of high molecular weight and composed of monomers which are trapped in a minium of conformational energy different from the one characterizing the native enzyme. This energy minimum is protected by a high activation energy of dissociation such that the aggregates are perfectly stable under nondenaturing conditions, and their degradation is provided only by applying strong denaturants, e.g., 6 M guanidine hydrochloride at neutral or acidic pH. The product of the slow redissolution process is the monomeric enzyme in its random configuration, which may be reactivated by diluting the denaturant under optimum conditions of reconstitution. The yield and the kinetics of reactivation of lactic dehydrogenase from pig skeletal muscle are not affected by the preceding aggregation-degradation cycle and are independent of different modes of aggregate formation (e.g., by renaturation at high enzyme concentration or heat aggregation). The kinetics of reactivation may be described by one single rate-determining bimolecular step with k2 = 3.9 x 10(4) M-1 s-1 at zero guanidine concentration. The reactivated enzyme consists of the native tetramer, characterized by enzymatic and physical properties identical with those observed for the enzyme in its initial native state.  相似文献   

20.
C J Kay  E W Lippay 《Biochemistry》1992,31(46):11376-11382
Kinetic and thermodynamic properties of yeast flavocytochrome b2 (EC 1.1.2.3) are modified by the product pyruvate, which binds to the flavosemiquinone (FSQ) form of the prosthetic flavin and decreases the thermodynamic driving force for electron transfer from FSQ to heme. Pyruvate inhibits flavocytochrome b2, but the catalytic competence of pyruvate-ligated FSQ in intramolecular electron transfer to heme is unclear; one kinetic study suggested pyruvate prevented this reaction [Tegoni, M, Janot J.-M., & Labeyrie, F. (1990) Eur. J. Biochem. 190, 329-342], while laser flash photolysis indicated pyruvate was essential [Walker, M. C., & Tollin, G. (1991) Biochemistry 30, 5546-5555]. To address this problem, wild-type (WT) and mutant (L36I) flavocytochromes b2 have been expressed in Escherichia coli. Both forms incorporated heme and FMN prosthetic groups and were catalytically active. The mutation L36I was a conservative substitution within the heme-binding crevice and was designed to alter the midpoint potential (Em) of the heme to alter the pyruvate-FSQ/heme equilibrium. Potentiometric titrations yielded Em values (pH 7.0, 25 degrees C) of +8 and -28 mV for WT and L36I forms, respectively. The FMN midpoint potentials in the absence of pyruvate (-58 mV, n = 2) were identical within experimental error in WT and L36I species and were also identical (+5 mV, n = 1) in the presence of pyruvate. These results indicated the absence of redox cooperativity between FMN and heme.(ABSTRACT TRUNCATED AT 250 WORDS)  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