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 46 毫秒
1.
L Powers  B M Kincaid 《Biochemistry》1989,28(10):4461-4468
Differences in the methods of analysis of X-ray absorption data used by Powers et al. [Powers, L., Blumberg, W. E., Chance, B., Barlow, C., Leigh, J., Jr., Smith, J., Yonetani, T., Vik, S., & Peisach, J. (1979) Biochim. Biophys. Acta 547, 520-538; Powers, L., Chance, B., Ching, Y., & Angiolillo, P. (1981) Biophys. J. 34, 465-498] and Scott et al. [Scott, R., Schwartz, J., & Cramer S. (1986) Biochemistry 25, 5546-5555] are clarified. In addition, we compare the X-ray absorption data and results for resting cytochrome c oxidase reported by both groups using the same analysis method and conclude apart from any assumptions that the data are not identical.  相似文献   

2.
In this work, we present a generalization of Zwanzig's protein unfolding analysis [Zwanzig, R., 1997. Two-state models of protein folding kinetics. Proc. Natl Acad. Sci. USA 94, 148-150; Zwanzig, R., 1995. Simple model of protein folding kinetics. Proc. Natl Acad. Sci. USA 92, 9801], in order to calculate the free energy change Delta(N)(D)F between the protein's native state N and its unfolded state D in a chemically induced denaturation. This Extended Zwanzig Model (EZM) is both based on an equilibrium statistical mechanics approach and the inclusion of experimental denaturation curves. It enables us to construct a suitable partition function Z and to derive an analytical formula for Delta(N)(D)F in terms of the number K of residues of the macromolecule, the average number nu of accessible states for each single amino acid and the concentration C(1/2) where the midpoint of the N<==>D transition occurs. The results of the EZM for proteins where chemical denaturation follows a sigmoidal-type profile, as it occurs for the case of the T70N human variant of lysozyme (PDB code: T70N) [Esposito, G., et al., 2003. J. Biol. Chem. 278, 25910-25918], can be splitted into two lines. First, EZM shows that for sigmoidal denaturation profiles, the internal degrees of freedom of the chain play an outstanding role in the stability of the native state. On the other hand, that under certain conditions DeltaF can be written as a quadratic polynomial on concentration C(1/2), i.e., DeltaF approximately aC(1/2)(2)+bC(1/2)+c, where a,b,c are constant coefficients directly linked to protein's size K and the averaged number of non-native conformations nu. Such functional form for DeltaF has been widely known to fit experimental measures in chemically induced protein denaturation [Yagi, M., et al., 2003. J. Biol. Chem. 278, 47009-47015; Asgeirsson, B., Guojonsdottir, K., 2006. Biochim. Biophys. Acta 1764, 190-198; Sharma, S., et al., 2006. Protein Pept. Lett. 13(4), 323-329; Salem, M., et al., 2006. Biochim. Biophys. Acta 1764(5), 903-912] so EZM can shed some light into the physical meaning of the experimental values for the a,b,c coefficients.  相似文献   

3.
The structures of membrane proteins are difficult to obtain by crystallography and may be altered by the detergents used in their extraction. X-ray absorption spectroscopy has been used to identify the structures of the copper atoms of the membrane-bound enzyme in mitochondria and in submitochondrial particles at respective concentrations of 100 and 200 micron of molar copper. To within the experimental error, the x-ray absorption spectra of the copper atoms of the membrane-bound and the Yonetani (Yonetani, T. (1961) J. Biol. Chem. 236, 1680-1688) purified oxidase are identical; all detectable shells of the active site indicate no alteration of structural parameters. Significant differences are found when compared to the Hartzell-Beinert (Hartzell, R. C., and Beinert, H. (1974) Biochim. Biophys. Acta 368, 318-338) preparation. Extended x-ray absorption fine structure technology is now adequate for the direct studies of membrane proteins in situ in their natural environment.  相似文献   

4.
A model for the water oxidation reaction in Photosystem II (PSII) is presented, based on an H atom abstraction mechanism. The model rationalises the S-state dependence of observed substrate water exchange kinetics [Biochim. Biophys. Acta 1503 (2001) 197] and assumes that H transfer occurs to an oxidised micro-oxo bridge oxygen on the S(3)-->S(4)-->S(0) transition. The model requires that only one Mn-pair and a Ca ion be directly involved in the substrate binding and catalytic function. The multiline signal observed in the S(0) state is shown to plausibly arise from such a system. A detailed molecular model of the three-metal site, assuming ligation by those residues identified by mutagenesis as Ca/Mn ligands is presented. This bears a resemblance to the dinuclear Mn site in Mn catalase and is generally consistent with the electron density map of cyanobacterial PSII recently presented [Proc. Natl. Acad. Sci. U. S. A. 100 (2003) 98].  相似文献   

5.
T Y Teng  H W Huang  G A Olah 《Biochemistry》1987,26(25):8066-8072
A previous extended X-ray absorption fine structure (EXAFS) study of photolyzed carboxymyoglobin (MbCO) [Chance, B., Fischetti, R., & Powers, L. (1983) Biochemistry 22, 3820-3829; Powers, L., Sessler, J. L., Woolery, G. L., & Chance, B. (1984) Biochemistry 23, 5519-5523] has provoked much discussion on the heme structure of the photoproduct (MbCO). The EXAFS interpretation that the Fe-CO distance increases by no more than 0.05 A following photodissociation has been regarded as inconsistent with optical, infrared, and magnetic susceptibility studies [Fiamingo, F. G., & Alben, J. O. (1985) Biochemistry 24, 7964-7970; Sassaroli, M., & Rousseau, D. L. (1986) J. Biol. Chem. 261, 16292-16294]. The present experiment was performed with well-characterized dry film samples in which MbCO molecules were embedded in a poly(vinyl alcohol) matrix [Teng, T. Y., & Huang, H. W. (1986) Biochim. Biophys. Acta 874, 13-18]. The sample had a high protein concentration (12 mM) to yield adequate EXAFS signals but was very thin (40 micron) so that complete photolysis could be easily achieved by a single flash from a xenon lamp. Although the electronic state of MbCO resembles that of deoxymyoglobin (deoxy-Mb), direct comparison of EXAFS spectra indicates that structurally MbCO is much closer to MbCO than to deoxy-Mb.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The oxidation of ferric cytochrome c peroxidase by hydrogen peroxide yields a product, compound ES [Yonetani, T., Schleyer, H., Chance, B., & Ehrenberg, A. (1967) in Hemes and Hemoproteins (Chance, B., Estabrook, R. W., & Yonetani, T., Eds.) p 293, Academic Press, New York], containing an oxyferryl heme and a protein free radical [Dolphin, D., Forman, A., Borg, D. C., Fajer, J., & Felton, R. H. (1971) Proc. Natl. Acad. Sci. U.S.A. 68, 614-618]. The same oxidant takes the ferrous form of the enzyme to a stable Fe(IV) peroxidase [Ho, P. S., Hoffman, B. M., Kang, C. H., & Margoliash, E. (1983) J. Biol. Chem. 258, 4356-4363]. It is 1 equiv more highly oxidized than the ferric protein, contains the oxyferryl heme, but leaves the radical site unoxidized. Addition of sodium fluoride to Fe(IV) peroxidase gives a product with an optical spectrum similar to that of the fluoride complex of the ferric enzyme. However, reductive titration and electron paramagnetic resonance (EPR) data demonstrate that the oxidizing equivalent has not been lost but rather transferred to the radical site. The EPR spectrum for the radical species in the presence of Fe(III) heme is identical with that of compound ES, indicating that the unusual characteristics of the radical EPR signal do not result from coupling to the heme site. By stopped-flow measurements, the oxidizing equivalent transfer process between heme and radical site is first order, with a rate constant of 0.115 s-1 at room temperature, which is independent of either ligand or protein concentration.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Phospholipid asymmetry is a fundamental feature of the plasma membrane of most eukaryotic cells and its regulation is linked to diverse physiological processes such as apoptosis and blood clotting [P. Williamson, R.A. Schlegel, Biochim. Biophys. Acta 1585 (2002) 53-63; R.F. Zwaal, A.J. Schroit, Blood 89 (1997) 1121-1132]. In addition, the phospholipid translocases (flippases) that are thought to establish asymmetry are also implicated in vesicle-mediated protein transport throughout the secretory and endocytic pathways [T.R. Graham, Trends Cell Biol. 14 (2004) 670-677]. However, the biochemical properties of phospholipid translocases in membranes of the Golgi complex and endosomes have received much less attention than translocases in the plasma membrane. We describe here a method for purifying yeast Golgi membranes and assaying an ATP-dependent phospholipid translocase activity in these membranes using fluorescent lipid analogues. This assay detects ATP-dependent translocation of labeled phosphatidylserine across late Golgi membranes, which requires the activity of a P-type ATPase called Drs2p [P. Natarajan, J. Wang, Z. Hua, T.R. Graham, Proc. Natl. Acad. Sci. USA 101 (2004) 10614-10619].  相似文献   

8.
The X-ray crystal structure of human soluble epoxide hydrolase (sEH) has been determined at 2.6 A resolution, revealing a domain-swapped quaternary structure identical to that observed for the murine enzyme [Argiriadi, M. A., Morisseau, C., Hammock, B. D., and Christianson, D. W. (1999) Proc. Natl. Acad. Sci. U.S.A. 96, 10637-10642]. As with the murine enzyme, the epoxide hydrolytic mechanism of the human enzyme proceeds through an alkyl-enzyme intermediate with Asp-333 in the C-terminal domain. The structure of the human sEH complex with N-cyclohexyl-N'-(iodophenyl)urea (CIU) has been determined at 2.35 A resolution. Tyr-381 and Tyr-465 donate hydrogen bonds to the alkylurea carbonyl group of CIU, consistent with the proposed roles of these residues as proton donors in the first step of catalysis. The N-terminal domain of mammalian sEH contains a 15 A deep cleft, but its biological function is unclear. Recent experiments demonstrate that the N-terminal domain of human sEH catalyzes the metal-dependent hydrolysis of phosphate esters [Cronin, A., Mowbray, S., Dürk, H., Homburg, S., Fleming, I., Fisslthaler, B., Oesch, F., and Arand, M. (2003) Proc. Natl. Acad. Sci. U.S.A. 100, 1552-1557; Newman, J. W., Morisseau, C., Harris, T. R., and Hammock, B. D. (2003) Proc. Natl. Acad. Sci. U.S.A. 100, 1558-1563]. The binding of Mg(2+)-HPO4(2-) to the N-terminal domain of human sEH in its CIU complex reveals structural features relevant to those of the enzyme-substrate complex in the phosphatase reaction.  相似文献   

9.
The 20-kDa fragment of myosin subfragment-1 heavy chain was cleaved with cyanogen bromide. Gel electrophoresis of the fragmented peptides indicated the presence of 20-, 18-, 16-, 14-, 12-, and 10-kDa peptides in addition to two peptides smaller than 10 kDa. The renaturation procedure of Muhlrad and Morales (Muhlrad, A., and Morales, M. (1984) Proc. Natl. Acad. Sci. U. S. A. 81, 1003-1007) was applied to the mixture of these peptides. The peptides larger than 10 kDa, which contain both the reactive SH1 and SH2 groups, were precipitated with F-actin by ultracentrifugation. The 10-kDa peptide was purified and was identified as p10 of Elzinga and Collins (Elzinga, M., and Collins, J. H. (1977) Proc. Natl. Acad. Sci. U. S. A. 74, 4281-4284). The renaturation procedure was applied to the purified 10-kDa peptide. The 10-kDa peptide was also precipitated with F-actin by ultracentrifugation. Affinity of the 10-kDa peptide for F-actin was determined with an increase of turbidity, and the apparent dissociation constant was 0.94 microM. Results are consistent with our proposition that a binding site for F-actin exists around the SH1 and SH2 groups of subfragment-1 (Katoh, T., Imae, S., and Morita, F. (1984) J. Biochem. 95, 447-454; Katoh, T., and Morita, F. (1984) J. Biochem. 96, 1223-1230).  相似文献   

10.
The finding that animal synthetases are complexes consisting of two polypeptide chains (Stoops, J.K., Arslanian, M.J., Oh, Y.H., Vanaman, T.C., and Wakil, S.J. (1975) Proc. Natl. Acad. Sci. U. S. A. 72, 1940-1944) led us to investigate their 4'-phosphopantetheine content. We have found that the chicken and rat synthetases contain 1.6 to 2.2 mol of 4'-phosphopantetheine per mol of the complex. The implications of this finding concerning the structure of the complex and the biosynthetic pathway of fatty acid synthesis are discussed.  相似文献   

11.
Intramolecular cross-linking of myosin subfragment 1 with bimane   总被引:2,自引:0,他引:2  
K Ue 《Biochemistry》1987,26(7):1889-1894
We previously showed that the fluorescent inter-thiol cross-linker dibromobimane (DBB) [Kosower, N. S., Kosower, E. M., Newton, G. L., & Ranney, H. M. (1979) Proc. Natl. Acad. Sci. U.S.A. 76, 3382-3386] cross-links two [50 and 20 kilodaltons (kDa)] of the three major fragments of myosin subfragment 1 (S-1); on intact S-1, DBB quenches tryptophans and inhibits all ATPases [Mornet, D., Ue, K., & Morales, M. F. (1985) Proc. Natl. Acad. Sci. U.S.A. 82, 1658-1662]. Here we characterize the modification chemically: DBB cross-links Cys-522 (50 kDa) with Cys-707 (20 kDa), thereby sealing a large preexisting heavy-chain loop containing important functionalities. Cross-linking rate is insensitive to nucleotides, but apparently sterically, either monobromobimane or DBB reduces Ca2+-ATPase to low, nonzero levels.  相似文献   

12.
Dilauroylglycerophosphocholine (C12:0PC)-induced vesiculation of platelet plasma membranes (Kobayashi, T., Okamoto, H., Yamada, J.-I., Setaka, M. and Kwan, T. (1984) Biochim. Biophys. Acta 778, 210-218; Kobayashi, T., Yamada, J.-I., Satoh, N., Setaka, M. and Kwan, T. (1985) Biochim. Biophys. Acta 817, 307-312) was inhibited by chlorpromazine. Preincubation of platelets with chlorpromazine was required for inhibition but incorporation of chlorpromazine into C12:0PC liposomes was not necessary for it, indicating that the observed inhibition of vesiculation was mainly due to the effect of chlorpromazine on platelets and not that on liposomes. The change in platelet membrane fluidity caused by chlorpromazine was not the cause of inhibition of vesiculation. The inhibition of vesiculation by various other calmodulin antagonists was also observed. The inhibitory activities of these calmodulin antagonists and chlorpromazine correspond very well to their abilities to bind to calmodulin. N-(6-Aminohexyl)-5-chloro-1-naphthalenesulfonamide (W-7) inhibited vesiculation but a structural analogue of it, N-(6-aminohexyl)-1-naphthalenesulfonamide (W-5), had no inhibitory activity. These results suggest the involvement of calmodulin in membrane vesiculation.  相似文献   

13.
Mice that lack the mitochondrial form of superoxide dismutase (SOD2) incur severe pathologies and mitochondrial deficiencies, including major depletion of complex II, as a consequence of buildup of endogenous reactive oxygen species (Melov, S., Coskun, P., Patel, M., Tuinstra, R., Cottrell, B., Jun, A. S., Zastawny, T. H., Dizdaroglu, M., Goodman, S. I., Huang, T. T., Miziorko, H., Epstein, C. J., and Wallace, D. C. (1999) Proc. Natl. Acad. Sci. U. S. A. 96, 846-851 and Li, Y., Huang, T. T., Carlson, E. J., Melov, S., Ursell, P. C., Olson, J. L., Noble, L. J., Yoshimura, M. P., Berger, C., Chan, P. H., Wallace, D. C., and Epstein, C. J. (1995) Nat. Genet. 11, 376-381). These problems can be greatly attenuated or rescued by synthetic antioxidant treatment, such as with the catalytic antioxidant EUK189 (Hinerfeld, D., Traini, M. D., Weinberger, R. P., Cochran, B., Doctrow, S. R., Harry, J., and Melov, S. (2004) J. Neurochem. 88, 657-667). We have used heart mitochondria from sod2 null mice to better understand mitochondrial reactive oxygen species production both in the absence of SOD2 and following in vivo antioxidant treatment. Isolated heart mitochondria from 5-day-old sod2 null animals respiring on the complex II substrate succinate exhibited statistically significant higher levels of mitochondrial O2* (157%, p < 0.01) but significantly less H2O2 (33%, p < 0.001) than wild type littermates. Treatment of sod2 nullizygous mice with EUK189 proportionately increased the levels of complex II and H2O2. Increased production of O2* resulting from complex II normalization had no effect on steady state levels due to the rapid conversion to H2O2, a process presumably aided by the presence of the EUK189, an SOD mimetic.  相似文献   

14.
Cells continuously probe their environment with membrane receptors, achieving subsecond adaptation of their behaviour [Diez, G., Gerisch, G., Anderson, K., Müller-Taubenberger, A. and Bretschneider, T. (2006) Subsecond reorganization of the actin network in cell motility and chemotaxis. Proc. Natl. Acad. Sci. USA 102, 7601-7606, Shamri, R., Grabovsky, V., Gauguet, J.M., Feigelson, S., Manevich, E., Kolanus, W., Robinson, M.K., Staunton, D.E., von Andrian, U.H. and Alon, R. (2005) Lymphocyte arrest requires instantaneous induction of an extended LFA-1 conformation mediated by endothelium-bound chemokines. Nat. Immunol. 6, 497-606, Jiang, G., Huang, A.H., Cai, Y., Tanase, M. and Sheetz, M.P. (2006) Rigidity sensing at the leading edge through alpha(V)beta(3) integrins and RPTPalpha. Biophys. J. 90, 1804-2006]. Recently, several receptors, including cadherins, were found to bind ligands with a lifetime of order of one second. Here we show at the single molecule level that homotypic C-cadherin association involves transient intermediates lasting less than a few tens of milliseconds. Further, these intermediates transitionned towards more stable states with a kinetic rate displaying exponential decrease with piconewton forces. These features enable cells to detect ligands or measure surrounding mechanical behaviour within a fraction of a second, much more rapidly than was previously thought.  相似文献   

15.
The enzymic degradation of a number of sphingolipids in the lysosomes is stimulated by small acid glycoproteins named activator proteins. We purified and sequenced a new protein, called component C, which seems to be related to sulfatide activator and to a recently described activator of glucosylceramidase (A1 activator) (Kleinschmidt, T., Christomanou, H. & Braunitzer, G. (1987) Biol. Chem. Hoppe-Seyler 368, 1571-1578). It consists of 78 amino acids and carries one carbohydrate chain at aparagine 20. Component C shows 21.5% sequence homology to sulfatide activator and 34.2% homology to A1 activator. Structural similarities between these three proteins have also been detected. Recently the cDNA sequence of the sulfatide activator precursor has been published (Dewji, N.N., Wenger, D.A. & O'Brien, J.S. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 8652-8656). We could align the protein sequences of sulfatide activator, A1 activator and component C with that of this large precursor protein. After minor corrections of the DNA sequence we obtained total fit. Thus it seems that three different proteins are derived from the sulfatide activator precursor by proteolytic processing. Possible processing sites were found on the precursor at sites adjacent to the N-termini and C-termini of the mature proteins. The processing of sulfatide activator was studied by Fujibayashi and Wenger (Fujibayashi, S. & Wenger, D.A. (1986) Biochim. Biophys. Acta 875, 554-562). Their data support our assumption that processing occurs by simultaneous cleavage at all possible sites.  相似文献   

16.
ATP-dependent activation and deactivation of retinal rod outer segment phosphodiesterase is affected by calcium [Kawamura, S. and Bownds, M. D., J. Gen. Physiol. 77:571-591(1981)]. Our data demonstrate that although calmodulin has been found in rod outer segments [Liu, Y. P. and Schwartz, H., Biochim. Biophys. Acta 526:186-193(1978); Kohnken, R. E. et al, J. Biol. Chem. 256:12517-12522(1981)], this protein is not involved in calcium-dependent phosphodiesterase activation at light levels at which calcium clearly affects this enzyme's activity. Furthermore, calmodulin does not mediate the calcium-dependent deactivation of phosphodiesterase.  相似文献   

17.
Calmodulin (CaM)-stimulated phosphatase in bovine brain or bovine lung CaM-binding protein fractions were fractionated on a heparin-Sepharose column into three activity peaks, designated in order of column three activity peaks, designated in order of column elution as the brain peak I (BPI), peak II (BPII), and peak III (BPIII) or the lung peak I (LPI), peak II (LPII), and peak III (LPIII) phosphatases, respectively. The pooled individual peak fractions were further purified on a fast protein liquid chromatography Superose 12 column. Analysis of the purified samples by sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed that they all contained major peptides corresponding to alpha and beta subunits of the brain CaM-stimulated phosphatase. The phosphatases had similar specific activities and were similarly stimulated by Ni2+, Mn2+, Mg2+ + Ca2+, and CaM. They showed differential reactivity on immunotransblots with an alpha subunit-specific monoclonal antibody VJ6, which reacted strongly toward BPI and weakly toward BPIII and LPI, but showed no reactivity toward BPII, LPII, and LPIII. Each of the alpha subunits of the purified phosphatases had a distinct V8 protease and chymotrypsin peptide map. The results suggest that both bovine brain and bovine lung contain multiple CaM-stimulated phosphatase isozymes. The suggestion of three mammalian brain CaM-stimulated phosphatase isozymes is in agreement with the results of recent molecular cloning studies (Kuno, T., Takeda, T., Hirai, M., Ito, A., Mukai, H., and Tanaka, C. (1989) Biochem. Biophys. Res. Commun. 165, 1352-1358; Guerini, D., and Klee, C.B. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 9183-9187; da Cruz e Silva, E. F., and Cohen, P. T. W. (1989) Biochim. Biophys. Acta 1009, 293-296). The successful purification of the individual isozymes may facilitate the elucidation of molecular basis and physiological significance of the isozymes.  相似文献   

18.
A model has been built of the amino-terminal domain of the intercellular adhesion molecule-1 (ICAM-1), the receptor for most human rhinovirus serotypes. The model was based on sequence and presumed structural homology to immunoglobulin constant domains. It fits well into the putative receptor attachment site, the canyon, on the human rhinovirus-14 (HRV14) surface in a manner consistent with most of the mutational data for ICAM-1 (Staunton, D. E., Dustin, M. L., Erickson, H. P., Springer, T. A. Cell, in press, 1989) and HRV14 (Colonno, R. J., Condra, J. H., Mizutani, S., Callahan, P. L., Davies, M. E., Murcko, M. A. Proc. Natl. Acad. Sci. U.S.A. 85: 5449-5453, 1988).  相似文献   

19.
20.
A comparison of DNA polymerase III core enzyme (McHenry, C. S., and Crow, W. (1979) J. Biol. Chem. 254, 1748-1753) prepared from wild type Escherichia coli and a strain harboring the mutator gene, mutD5 (Degnen, G. E., and Cox, E. C. (1974) J. Bacteriol. 17, 477-487) has revealed several differences in their properties. Among these are alterations in the heat stability, divalent cation requirement, pH optimum, 3'----5'-single strand exonuclease activity, and DNA-dependent conversion of a deoxynucleoside triphosphate to its corresponding monophosphate ("turnover"). The decrease in the 3'-single strand exonuclease and turnover indicate a defect in the editing function of the mutD strain, which is at least in part responsible for the high spontaneous mutation rate in mutD. Transformation of mutD by a hybrid plasmid, pRD3, constructed from an EcoRI restriction fragment of E. coli and pBR322, cures mutD of its abnormally high mutation rate, and simultaneously restores its 3'-exonuclease activity. These observations are consistent with the notion that the mutD gene product is a subunit of DNA polymerase III, and it either contains the catalytic site for the 3'-exonuclease or modulates its activity. From a consideration of the known molecular weights of the subunits in DNA polymerase III core (McHenry C. S., and Crow, W. (1979) J. Biol. Chem. 254, 1748-1753) the molecular weights of the two proteins translated in maxicells transformed with pRD3, and from a comparison of our results with those obtained with the mutator dnaQ (Horiuchi, T., Maki, H., Maruyama, M., and Sekiguchi, M. (1981) Proc. Natl. Acad. Sci. U. S. A. 78, 3770-3774) and the work of Cox and Horner (Cox, E. C., and Horner, D. L. (1983) Proc. Natl. Acad. Sci. U. S. A. 80, 2295-2299) as well as Echols et al. (Echols, H., Lu, C., and Burgers, P. M. J. (1983) Proc. Natl. Acad. Sci. U. S. A. 80, 2189-2192) we tentatively assign the mutD gene product to the epsilon subunit of DNA polymerase III.  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