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 0 毫秒
1.
alpha-Lactalbumin (alpha-LA) is a calcium binding protein that also binds Mn(II), lanthanide ions, A1(III), Zn(II), Co(II). The structural implications of cation binding were studied by high-resolution proton (200 MHz) NMR and photochemically induced dynamic nuclear polarization (CIDNP) spectroscopy. Marked changes were observed in the NMR spectra of the apoprotein upon addition of a stoichiometric amount of calcium to yield Ca(II)-alpha-LA, manifested particularly in ring current shifted aliphatic peaks and in several shifts in the aromatic region, all of which were under slow exchange conditions. The CIDNP results showed that two surface-accessible tyrosine residues, assigned as Tyr-18 and -36, became inaccessible to the solvent upon addition of 1:1 Ca(II) to apo-alpha-lactalbumin, while Tyr-103 and Trp-104 remained completely accessible in both conformers. The proton NMR spectra of apo-alpha-LA and A1(III)-alpha-LA were extremely similar, which was also consistent with intrinsic fluorescence results [Murakami, K., & Berliner, L. J. (1983) Biochemistry 22, 3370-3374]. The paramagnetic cation Mn(II) bound to the strong calcium binding site on apo-alpha-LA but also to the weak secondary Ca(II) binding site(s) on Ca(II)-alpha-LA. It was also found that Co(II) bound to some secondary sites on Ca(II)-alpha-LA that overlapped the weak calcium site. All of the lanthanide shift reagents [Pr(III), Eu(III), Tb(III), Dy(III), Tm(III), Yb(III)] bound under slow exchange conditions; their relative affinities for apo-alpha-lactalbumin from competitive binding experiments were Dy(III), Tb(III), and Pr(III) greater than Ca(II) greater than Yb(III).  相似文献   

2.
Recently, solid-state NMR spectroscopy became a viable method to investigate photosynthetic reaction centres (RCs) on the atomic level. To study the electronic structure of the radical cation state of the RC, occurring after the electron emission, solid-state NMR using an illumination set-up can be exploited. This paper describes the illumination set-up we designed for a standard Bruker wide-bore MAS NMR probe. In addition we demonstrate its application to get information from the active site in photosynthetic reaction centres of Rhodobacter sphaeroides R-26 by photochemically induced dynamic nuclear polarization (photo-CIDNP). Solid-state NMR spectra of natural abundance 13C in detergent solubilized quinone depleted photosynthetic reaction centres under continuous illumination showed exceptionally strong nuclear spin polarization in NMR lines. Both enhanced-absorptive and emissive polarization were seen in the carbon spectrum which could be assigned to a bacteriochlorophyll a (BChl a) cofactor, presumably the special pair BChl a. The sign and intensities of the 13C NMR signals provide information about the electron spin density distribution of the transiently formed radical P.+ on the atomic level.  相似文献   

3.
Controlled proteolysis of epidermal growth factor from the mouse leads to fragments of mouse epidermal growth factor containing residues 1-48 and 1-45. The COOH-terminal pentapeptide appears to play a crucial role in determining the hydrophobic interactions between the hormone and the stationary phase during gel chromatography on TSK-125 gel. Proton NMR studies indicate that the overall structure of mouse epidermal growth factor is retained in the protein devoid of the COOH-terminal pentapeptide, while subsequent cleavage of the peptide bond between Arg-45 and Asp-46 starts to perturb the proton resonances most characteristic of the tertiary structure of the hormone, especially those from the aromatic ring protons of Tyr-37. Consequently, photochemically induced dynamic nuclear polarization experiments show an increased exposure of Tyr-37 in the fragment of mouse epidermal growth factor containing residues 1-48. Nuclear Overhauser data suggest that structural changes do occur on fragmentation but seem to be localized in the tiered-beta-sheet domain which contains Tyr-37.  相似文献   

4.
M M Snel  R Kaptein  B de Kruijff 《Biochemistry》1991,30(14):3387-3395
The topology of apocytochrome c, the heme-free precursor of the mitochondrial protein cytochrome c, was investigated in a lipid-associated form. For this purpose photochemically induced dynamic nuclear polarization 1H nuclear magnetic resonance (CIDNP 1H NMR) spectroscopy and quenching of tryptophan and tyrosine fluorescence by acrylamide were applied to an apocytochrome c-sodium dodecyl sulfate (SDS) micellar system. A pH titration of the chemical shifts of the histidine C2 proton resonances of apocytochrome c, using conventional 1H NMR, yielded pK(a)'s of 5.9 +/- 0.1 and 6.2 +/- 0.1, which were assigned to histidine-18 and -33 and histidine-26, respectively. In the presence of SDS micelles an average pK(a) of 8.1 +/- 0.1 was obtained for all histidine C2 protons. Photo-CIDNP enhancements of the histidine, tryptophan, and tyrosine residues, contained in the intact apocytochrome c and in chemically and enzymatically prepared fragments of the precursor, were reduced in the presence of SDS micelles. Similarly, the quenching of the tryptophan fluorescence of the polypeptides by acrylamide was diminished in the presence of SDS. These results indicate the aromatic residues studied are localized in the interface of the SDS micelle.  相似文献   

5.
Photochemically induced dynamic nuclear polarization (CIDNP)-1H-NMR spectroscopy has been used to study the interaction of the protein hormone epidermal growth factor (EGF) with micelles of sodium dodecyl sulfate (SDS) and dodecylphosphorylcholine (DPC). Conventional 1H-NMR spectra show that most protein resonances remain unperturbed when micelles are added to solution, which argues that the overall protein conformation is maintained in the presence of SDS or DPC at the concentrations used. Photo-CIDNP enhancements of resonances assigned to aromatic side chains of residues at the COOH terminus and beta-sheet regions of murine EGF (i.e. Trp-49, Trp-50, and Tyr-37) are considerably reduced in the presence of micelles, while resonances of aromatic side chains of residues found elsewhere on the protein surface are mostly unaffected. This suggests that the primary interaction between murine EGF and the micelle occurs at the micelle-bulk solvent interface. The overall negatively charged surface of SDS micelles tends to induce a stronger interaction with the protein compared to the zwitterionic DPC micelles, probably due to electrostatic interactions. Cleavage of the COOH-terminal pentapeptide containing both tryptophan residues enhances the already present, but weak, interaction with Tyr-10 and attenuates it with Tyr-37. A similar interaction pattern is found with rat EGF suggesting that at least concerning these two species of EGF the interaction is somewhat specific and conserved. A simple mass-action model for protein-micelle interaction is also presented.  相似文献   

6.
Lacking the extraordinary thermal stability of its metal-bound forms, apo-alpha-parvalbumin from rat muscle assumes two distinct conformations in aqueous solution. At 25 degrees C, its highly structured form predominates (Keq = 5.7; delta G degree = -4.3 kJ X mol-1); as deduced from both 1H NMR and circular dichroism (CD) spectroscopy, this conformation is exceedingly similar to those of its Mg(II)-, Ca(II)-, and Lu(III)-bound forms. The temperature dependences of several well-resolved aromatic and upfield-shifted methyl 1H NMR resonances and several CD bands indicate that the native, highly helical structure of rat apo-alpha-parvalbumin is unfolded by a concerted mechanism, showing no indication of partially structured intermediates. The melting temperature, TM, of rat apo-alpha-parvalbumin is 35 +/- 0.5 degrees C as calculated by both spectroscopic techniques. By 45 degrees C, rat apo-alpha-parvalbumin unfolds entirely, losing the tertiary structure that characterizes its folded form: not only are the ring-current-shifted aromatic and methyl 1H NMR resonances leveled, but the 262- and 269-nm CD bands are also severely reduced. As judged by the decrease in the negative ellipticity of the 222-nm CD band, this less-structured form of rat apo-alpha-parvalbumin shows an approximate 50% loss in apparent alpha-helical content compared to its folded state. Several changes in the 1H NMR spectrum of rat apo-alpha-parvalbumin were exceptionally informative probes of the specific conformational changes that accompany metal ion binding and metal ion exchange. In particular, the line intensities of the ortho proton resonance of Phe-47, the unassigned downfield-shifted alpha-CH resonances from the beta-sheet contacts between the metal-binding loops, the C2H resonance of His-48, and the epsilon-CH3 resonance of an unassigned Met residue were monitored as a function of added metal to determine the stability constants of several metal ion-parvalbumin complexes. We conclude that Mg(II) binds to the CD and EF sites independently, its affinity for the EF site being almost twice that for the CD site. Mg(II)----Ca(II) exchange showed that the CD-site Mg(II) is displaced first, in contrast to Lu(III)'s preferential displacement of the EF-site Ca(II) as determined from the Ca(II)----Lu(III) exchange experiments.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
A series of modified parvalbumins, differing only in length of alpha-helix F at the C-terminus, was prepared by carboxypeptidase-mediated digestions of the beta-lineage parvalbumin (pI = 4.25) from carp (N; 108 residues). Removal of Ala-108 to form the N-1 derivative (des-Ala108,Lys107-parvalbumin) only slightly alters the protein's ability to chelate Ca(II) or lanthanides(III). Analysis of the kinetics of their Yb(III) off-rates by optical stopped-flow techniques, determination of their Lu(III)-binding constants by high-resolution 1H NMR methods, and inspection of their solution structures by Yb(III)-shifted 1H NMR techniques indicate N-1 and N-2 are very similar to N (0.1-0.2 M KCl; pH 6-7; 23-55 degrees C). However, removal of the next one or two residues, Val-106 or Val-106/Leu-105, to generate the N-3 and N-4 derivatives severely alters the metal ion binding characteristics of the protein. Although two Yb(III) off-rates are observed for N-3, both are faster than that for the unmodified protein: kCD by a factor of 2 and kEF by a factor of 2200. Removal of Ala-104 and Ala-104/Thr-103 to give a mixture of N-5 and N-6 derivatives eliminates the slow-release site altogether, the single observable koff being 20-30 times faster than release of Yb(III) from the CD site of native parvalbumin. Removal of the C-terminal alpha-helix by digestion through Phe-102 to give N-7 destabilizes the entire protein structure as judged both by the random-coil appearance of its 1H NMR spectrum and by its aberrant kinetics. Although one abnormally fast koff is still observed at micromolar concentrations, Ln(III) chelation tends to precipitate N-7 at higher parvalbumin concentrations (1-3 mM). In contrast to the critical instability of the N-3 through N-7 derivatives, the remarkable stability of the N-1 and N-2 forms of carp parvalbumin may be attributed to the maintainance of two key structural features: an ion pair bond between the negatively charged C-terminal carboxyl function and the protonated epsilon-NH3+ of Lys-27 and hydrophobic interactions of the inner side of helix F with residues in the protein's core.  相似文献   

8.
Human endothelin-1 (ET-1) is a potent cardiovascular bioactive peptide. Its activity is based on the C-terminal residues, e.g., Trp 21 in particular. Recently, we reported an NMR solution structure of ET-1, which has a C-terminal hydrophobic core around Tyr 13. This C-terminal conformation does not agree with a previously reported X-ray crystal structure. To clarify the discrepancy, we performed photo-CIDNP NMR in combination with MALDI-TOF MS. The photo-CIDNP results revealed that the Tyr 13 aromatic ring is concealed in a hydrophobic interaction. MALDI-TOF MS experiments showed this is an intramolecular interaction in monomeric form, which is also supported by sedimentation analysis and two-dimensional NMR cross-peak line shapes. Thus, we confirmed the intramolecular hydrophobic core around Tyr 13 in aqueous solution, which agrees with the solution structure. The C-terminal conformational discrepancy between the solution and crystal was caused by the intermolecular hydrogen bond between Tyr 13 of one molecule and Asp 8 of the other in a dimer-like formation of crystalline ET-1. On the other hand, we indicated that endothelin-3, another isoform of the endothelin, has an apparent self-association equilibrium under the same condition in which three tyrosines participate.  相似文献   

9.
R E Klevit  E B Waygood 《Biochemistry》1986,25(23):7774-7781
Sequence-specific resonance assignments of the 1H NMR spectrum of the 85-residue histidine-containing phosphocarrier protein (HPr) are complete [Klevit, R. E., Drobny, G. P., & Waygood, E. B. (1986) Biochemistry (first paper of three in this issue)]. Additional side-chain assignments have been made with long-range coherence transfer experiments [Klevit, R. E., & Drobny, G. P. (1986) Biochemistry (second paper of three in this issue)]. In this paper, the NMR assignments were used to determine the secondary structure and the tertiary folding of HPr in solution. The secondary structural elements of the protein were determined by visual inspection of the pattern of nearest-neighbor nuclear Overhauser effects (NOEs) and the presence of persistent amide resonances. Escherichia coli HPr consists of four beta-strands, three alpha-helices, four reverse turns, and several regions of extended backbone structure. Long-range NOEs, especially among side-chain protons, were used to determine the tertiary structure of the protein by use of the secondary structural components. The four beta-strands form a single antiparallel beta-pleated sheet. The hydrophobic faces of the alpha-helices interact to form a hydrophobic core and sit above the hydrophobic face of the beta-sheet, forming an open-face beta-sheet sandwich structure. The active site histidine, His-15, is on a short kinked segment of backbone that is accessible to the solvent. The positively charged phosphorylation site (His-15 and Arg-17) interacts with the negatively charged carboxyl terminus of the protein (Glu-85).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Degradation intermediates of o-, m- and p-cresols extracted from resting cells of Pseudomonas sp. CP4, a potent cresol- and phenol-degrading laboratory isolate, were analysed by using 1H NMR spectroscopy at 270 MHz. Ortho-, meta- and para-cresols were found to be degraded to 2-methyl-4-oxalocrotonate. 3-Methylcatechol from o-cresol was degraded further to 2-ketohex-cis-4-enoate, 4-methylcatechol from m- and p-cresol was degraded to 2-ketohex-cis-4-enoate. Also 2-ketopent-4-enoate was found to be formed from p-cresol. Formation of 2-methyl-4-oxalocrotonate was envisaged as taking place from 5-hydroxy-2-methylmuconic semialdehyde, the ring-cleavage product of 4-methylresorcinol, a possible product by hydroxylation of o-cresol along with 3-methylcatechol. This is a deviation from the hitherto known pathways of o-cresol degradation. Based on these observations, pathways for the degradation of all three isomers of cresol are proposed.  相似文献   

11.
The role of lipid diversity in biomembranes is one of the major unsolved problems in biochemistry. One parameter of possible importance is the mean cross-sectional area occupied per lipid molecule, which may be related to formation of nonbilayer structures and membrane protein function. We have used 2H NMR spectroscopy to compare the properties of 1,2-diperdeuteriopalmitoyl-sn-glycero-3-phosphoethanolamine (DPPE-d62) and 1,2-diperdeuteriopalmitoyl-sn-glycero-3-phosphocholine (DPPC-d62) in the L alpha phase. We find that DPPE has greater segmental order than DPPC, and that this increase in order is related to the smaller area per acyl chain found for DPPE. Values of the mean cross-sectional chain area are calculated using a simple diamond lattice model for the acyl chain configurational statistics, together with dilatometry data. The results obtained for the mean area per molecule are comparable with those from low angle x-ray diffraction studies.  相似文献   

12.
The 1H-NMR spectra of the two Escherichia coli basic, low-Mr (approximately equal to 9000) DNA-binding proteins NS1 and NS2 and of their native complex NS were studied at 400 MHz and a number of resonances and resonance peaks were assigned. As in the case of some eukaryotic histones, the presence of a large number of high-field perturbed Phe resonances, several shielded and deshielded methyl resonances and backbone NH protons quite inaccessible to the solvent clearly indicate the existence of extensive tertiary and, even more so, quaternary structures involving hydrophobic interactions. These structures are lost upon heating, but readily reform upon cooling. Spectral differences between NS1, NS2 and NS and the greater thermal stability of NS indicate that molecules of the heterologous subunits (NS1 and NS2) aggregate (dimerize) preferentially in comparison to the self-aggregation of the homologous subunits. Unlike those of the eukaryotic histones, the tertiary and quaternary structures of NS are insensitive to extensive variations of the ionic strength.  相似文献   

13.
As deduced from its 1H NMR spectrum, oncomodulin's solution conformation is very similar to the tertiary structure of other single domain 2-site calcium-binding proteins of the troponin C class. Despite its extensive amino acid sequence homology with parvalbumins, however, oncomodulin differs significantly from these proteins in its Ca(II)----Ln(III) exchange characteristics. Although the relative affinity of Lu(III) for the EF site of Ca2-oncomodulin was normal, beta Lu:EF/beta Ca:EF being 175 +/- 15, displacement of Ca(II) from the CD site was not favored, beta Lu:CD/beta Ca:CD being 1.2 +/- 0.1. Lineshape analyses of several 1H NMR resonances generated by the Lu(III) titration of Ca2-oncomodulin indicated that Ca(II)----Ln(III) exchange at the CD site was 15-20 s-1, approximately 100 times faster than exchange at the CD site of parvalbumins. Analyses of the distribution of metal-bound oncomodulin species showed that Ca(II)----Lu(III) exchange was cooperative, the coefficient of cooperativity being estimated as 5 +/- 1. The kinetics of the release of Yb(III) from oncomodulin as measured by optical stopped-flow techniques corroborated the observed cooperativity in metal binding; the off-rate constant of Yb(III) from the EF site of Yb2-oncomodulin was 0.0036 s-1, approximately 19 times slower than the release of Yb(III) from the EF site of Ca1Yb1-oncomodulin. We attribute part of the reduced preference of small Ln(III)s for the CD site of oncomodulin to a combination of this site's inherent incompressibility (Williams, T.C., Corson, D.C. & Sykes, B.D. (1984) J. Am. Chem. Soc. 106, 5698-5702) and the Glu----Asp substitution at sequence position 59, the residue which chelates metal at the -X coordination position. Like the CD site in oncomodulin, site III in troponin C has not only a lower affinity for calcium relative to the CD site of parvalbumins but also aspartic acid at its -X position; a water molecule bridges the gap between bound metal and the carboxyl group of the relatively short side chain of Asp-114 (Herzberg, O. & James, M. N. G. (1985) Biochemistry 24, 5298-5302). Hence, we suggest that Asp-59 in oncomodulin binds metal only indirectly through an intervening water molecule, a proposal which is consistent with the CD site's reduced affinity for ions the size of Ca(II) or smaller.  相似文献   

14.
15.
Possible pitfalls in mapping studies utilizing the nuclear relaxation rates induced by paramagnetic probes are pointed out. In cases in which a distance is sought between a paramagnetic ion and a small molecule (e.g. substrate, inhibitor, etc.), both bound non-covalently to a macromolecule, heterogeneity in the system with respect to the binding of either of them may result in ambiguous conclusions. It is shown that the trypsin-gadolinium (III)-inhibitor system is heterogeneous, as revealed in the dependence of the water and inhibitor proton line-widths upon both the Gd3+ and the enzyme concentrations and in the effects of added Ca2+ on the line-widths. The results imply that in published work (Abbott et al. (1975) Biochemistry 14, 4935) the distance from a weak rather than from the strong metal ion binding site of trypsin (EC 3.4.21.4) may have been determined.  相似文献   

16.
J N Breg  R Boelens  A V George  R Kaptein 《Biochemistry》1989,28(25):9826-9833
The Arc repressor of bacteriophage P22 is a DNA binding protein that does not belong to any of the known classes of such proteins. We have undertaken a 1H NMR study of the protein with the aim of elucidating its three-dimensional structure in solution and its mode of binding of operator DNA. Here we present the 1H nuclear magnetic resonance (NMR) assignments of all backbone protons and most of the side-chain protons of Arc repressor. Elements of secondary structure have been identified on the basis of networks of characteristic sequential and medium-range nuclear Overhauser enhancements (NOEs). Two alpha-helical regions have been found in the peptide regions 16-29 and 35-45. The ends of the helices could not yet be firmly established and could extend to residue 31 for the first helix and to residue 49 for the second. Immediately before the first helix, between residues 8 and 14, a region is present with beta-sheet characteristics dominated by a close proximity of the alpha-protons of residues 9 and 13. Because of the dimeric nature of the protein there are still two possible ways in which the NOEs in the beta-sheet region can be interpreted. If the NOEs are intramonomer, this requires a tight turn involving residues 10-12. Alternatively, if the NOEs are intermonomer, then and antiparallel beta-sheet would be implicated comprising two strands of different Arc monomers. While the data presently do not allow an unambiguous choice between these two possibilities, some evidence is discussed that favors the latter (beta-sheet between monomers).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Neuropeptide gamma belongs to tachykinin families which have a common C-terminal amino acid sequence (Phe-X-Leu-Met-NH2) and which induce various biological responses including salivation, hypotension, and contraction of gastrointestinal, respiratory, and urinary smooth muscle. In the present study, we present the solution structures of neuropeptide gamma (NPgamma) from gold fish (G-NPgamma) and mammalian NPgamma (M-NPgamma), as determined by nuclear magnetic resonance (NMR) spectroscopy in 50% trifluoroethanol (TFE)/water (1 : 1, v/v) solution and 200 mm sodium dodecyl sulfate (SDS) micelles. In aqueous TFE solution, G-NPgamma has a alpha-helical conformation in the region of His12-Met21 and a short helix in the N-terminal region, and has a beta-turn from Arg9 to Arg11 in between. In aqueous TFE solution, M-NPgamma also has alpha-helical conformations both in the C-terminal region and the N-terminal region and a beta-turn from His9 to Arg11 in between. In SDS micelle, the structure of G-NPgamma contains a stable alpha-helix from His12 to Met21 and a beta-turn from Arg9 to Arg11, while M-NPgamma has a short helix from Ser16 to Met21. The region from His12 to Met21 corresponds to the amino acid sequence of neurokinin A. Neuropeptide gamma may act as a precursor of neurokinin A and the post-translational processing of this peptide involves the enzymatic attack of the basic beta-turn region from residue 9 to residue 11 in the middle. From our relaxation study, it could be suggested that in fish system G-NPgamma induces the biological actions corresponding to those of substance P in mammalian system. The structures of G-NPgamma and M-NPgamma contain alpha-helical structures at the C-terminus and this helix seems to promote the affinity for NK1 and/or NK2 receptor.  相似文献   

18.
Photochemically induced dynamic nuclear polarization has been observed in reaction centres of the green sulphur bacterium Chlorobium tepidum by (13)C magic-angle spinning solid-state NMR under continuous illumination with white light. An almost complete set of chemical shifts of the aromatic ring carbons of a BChl a molecule has been obtained. All light-induced (13)C NMR signals appear to be emissive, which is similar to the pattern observed in the reaction centers of plant photosystem I and purple bacterial reaction centres of Rhodobacter sphaeroides wild type. The donor in RCs of green sulfur bacteria clearly differs from the substantially asymmetric special pair of purple bacteria and appears to be similar to the more symmetric donor of photosystem I.  相似文献   

19.
V J Basus  R M Scheek 《Biochemistry》1988,27(8):2772-2775
Complete sequence-specific assignments of the 1H NMR spectrum of bungarotoxin were reported in the previous paper [Basus, V.J., Billeter, M., Love, R.A., Stroud, R.M., & Kuntz, I.D. (1988) Biochemistry (first paper of three in this issue)]. The assignment was significantly aided by the use of the homonuclear Hartman-Hahn relayed coherence transfer nuclear Overhauser enhancement spectroscopy experiment (HRNOESY) which we present here, as a modification of relayed coherence transfer nuclear Overhauser enhancement spectroscopy (relayed NOESY) [Wagner, G. (1984) J. Magn. Reson. 57, 497]. As shown here, HRNOESY resolves problems of proton resonance overlap especially in extended chain conformations as found in beta-sheets.  相似文献   

20.
Binding of carbamoyl phosphate to Escherichia coli ornithine transcarbamoylase and its relation to turnover have been examined as a function of pH under steady-state conditions. The pH profile of the dissociation constant of carbamoyl phosphate (Kiacp) shows that the affinity of the substrate increases as pH decreases. Two ionizing groups are involved in carbamoyl phosphate binding. Protonation of an enzymic group with pKa 9.6 results in productive binding of the substrate with a moderate affinity of Kiacp approximately 30 microM. Protonation of a second group further enhances binding by roughly another order of magnitude. This ionization occurs with a pKa that shifts from less than 6 in the free enzyme to 7.3 in the binary complex. However, tighter binding of carbamoyl phosphate due to this ionization does not contribute to catalysis. The turnover rate (kcat) of the enzyme diminishes in the acidic pH range and is governed by an ionization with a pKa of 7.2. Both the catalytic pKa of 7.2 and the productive binding pKa of 9.6 appear in the pH profile of kcat/KMcp. Together with earlier kinetic results (Kuo, L. C., Herzberg, W., and Lipscomb, W. N. (1985) Biochemistry 24, 4754-4761), these data suggest that the step which modulates kcat may occur prior to the binding of the second substrate L-ornithine.  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