首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Memapsin 2 is the protease known as beta-secretase whose action on beta-amyloid precursor protein leads to the production of the beta-amyloid (Abeta) peptide. Since the accumulation of Abeta in the brain is a key event in the pathogenesis of Alzheimer's disease, memapsin 2 is an important target for the design of inhibitory drugs. Here we describe the residue preference for the subsites of memapsin 2. The relative k(cat)/K(M) values of residues in each of the eight subsites were determined by the relative initial cleavage rates of substrate mixtures as quantified by MALDI-TOF mass spectrometry. We found that each subsite can accommodate multiple residues. The S(1) subsite is the most stringent, preferring residues in the order of Leu > Phe > Met > Tyr. The preferences of other subsites are the following: S(2), Asp > Asn > Met; S(3), Ile > Val > Leu; S(4), Glu > Gln > Asp; S(1)', Met > Glu > Gln > Ala; S(2)', Val > Ile > Ala; S(3)', Leu > Trp > Ala; S(4)', Asp > Glu > Trp. In general, S subsites are more specific than the S' subsites. A peptide comprising the eight most favored residues (Glu-Ile-Asp-Leu-Met-Val-Leu-Asp) was found to be hydrolyzed with the highest k(cat)/K(M) value so far observed for memapsin 2. Residue preferences at four subsites were also studied by binding of memapsin 2 to a combinatorial inhibitor library. From 10 tight binding inhibitors, the consensus preferences were as follows: S(2), Asp and Glu; S(3), Leu and Ile; S(2)', Val; and S(3)', Glu and Gln. An inhibitor, OM00-3, Glu-Leu-Asp-LeuAla-Val-Glu-Phe (where the asterisk represents the hydroxyethylene tansition-state isostere), designed from the consensus residues, was found to be the most potent inhibitor of memapsin 2 so far reported (K(i) of 3.1 x 10(-10) M). A molecular model of OM00-3 binding to memapsin 2 revealed critical improvement of the interactions between inhibitor side chains with enzyme over a previous inhibitor, OM99-2 [Ghosh, A. K., et al. (2000) J. Am. Chem. Soc. 14, 3522-3523].  相似文献   

2.
The residues P3, P2, P1, and P1' of a peptide corresponding to the matrix/capsid protein junction in the HIV-1 gag protein (Ser-Gln-Asn-Tyr-Pro-Ile-Val) were systematically replaced and the effect of these single amino acid substitutions on the hydrolysis of each peptide by HIV-1 proteinase was studied. Subsites S1 and S1' of the enzyme showed explicit preference for hydrophobic moieties, but beta-branched amino acids and proline are not tolerated in S1. The S2 subsite shows a preference for small polar and apolar amino acids; it may be occupied by Asn, Asp, Glu, Cys, Ala, or Val, other substitutions, especially by Gln and Ser, prevent hydrolysis of the peptides. In subsite S3 all amino acids except proline can be accommodated.  相似文献   

3.
We have previously reported the asymmetric synthesis of (alpha-aminoalkyl) diphenylphosphonate and phosphinate derivatives designed as inhibitors of chymotrypsin- and elastase-like proteases. This paper reports the first kinetic evaluation of individual epimers of the (alpha-aminoalkyl) diphenylphosphonates as inactivators of chymotrypsin, cathepsin G and neutrophil elastase (HNE). Results show that the (R)-epimers consistently function as more potent irreversible inactivators of their respective target proteases than the corresponding (S)-epimers. Additionally, phosphinate analogues were found to be consistently superior to their diphenylphosphonate counterparts. For example, Cbz. Phe(P)(OPh)-(CH(2))(2)-CO(2)Et inactivates cathepsin G approximately 45-fold more rapidly (k(i)/K(i) = 1.2 x 10(5) M(-1). min(-1)) than the analogous Cbz.Phe(P)(OPh)(2) (2.6 x 10(3) M(-1). min(-1)). Similarly, Cbz.Val(P)(OPh)-(CH(2))(2)-CO(2)Et was found to inactivate HNE some 3-fold more efficiently than Cbz.Val(P)(OPh)(2) (6.5 x 10(3) and 2.0 x 10(3) M(-1). min(-1), respectively).  相似文献   

4.
To test the hypothesis that Glu202, adjacent to the His201 residue that participates in the coordination of Zn(2+) in matrix metalloproteinase-3 (MMP-3 or stromelysin-1), plays a role in its enzymatic activity it was substituted with Ala, Lys or Asp by site-specific mutagenesis. Wild-type proMMP-3, proMMP-3(E202A), proMMP-3(E202K) and proMMP-3(E202D) were expressed in Escherichia coli and purified to apparent homogeneity. Whereas 33-kDa wild-type proMMP-3 (consisting of the propeptide and catalytic domains) was quantitatively converted to 24-kDa active MMP-3 by treatment with p-aminophenyl-mercuric acetate (APMA), proMMP-3(E202A) and proMMP-3 (E202K) were fully resistant to APMA and proMMP-3 (E202D) was quantitatively converted into a 14-kDa species. In contrast, treatment with plasmin quantitatively converted the wild-type and the three mutant proMMP-3 moieties into the corresponding 24-kDa MMP-3 moieties. Biospecific interaction analysis revealed comparable affinity for binding to plasminogen of wild-type and mutant proMMP-3 (K(a) of 2.6-6.3 x 10(6) M(-1)) or MMP-3 (K(a) of 33-58 x 10(6) M(-1)) moieties. The affinity for binding to single-chain urokinase-type plasminogen activator (scu-PA) was also similar for wild-type and mutant proMMP-3 (K(a) of 5.0-6.9 x 10(6) M(-1)) or MMP-3 (K(a) of 37-72 x 10(6) M(-1)) moieties. However, MMP-3(E202A) and MMP-3(E202K) did not hydrolyze plasminogen whereas MMP-3(E202D) showed an activity of 20--30% of wild-type MMP-3. All three mutants were inactive towards scu-PA under conditions where this was quantitatively cleaved by wild-type MMP-3. Furthermore, MMP-3(E202A) and MMP-3(E202K) were inactive toward a fluorogenic substrate and MMP-3 (E202D) displayed about 15% of the activity of wild-type MMP-3. Taken together, these data suggest that Glu202 plays a crucial role in the enzymatic activity of MMP-3.  相似文献   

5.
Truncation of a peptide substrate in the N-terminus and replacement of its scissile amide bond with a non-cleavable reduced bond results in a potent inhibitor of HIV-1 protease. A series of such inhibitors has been synthesized, and S2-S3' subsites of the protease binding cleft mapped. The S2 pocket requires bulky Boc or PIV groups, large aromatic Phe residues are preferred in P1 and P1' and Glu in P2'. The S3' pocket prefers Phe over small Ala or Val. Introduction of a Glu residue into the P2' position yields a tight-binding inhibitor of HIV-1 protease, Boc-Phe-[CH2-NH]-Phe-Glu-Phe-OMe, with a subnanomolar inhibition constant. The relevant peptide derived from the same amino acid sequence binds to the protease with a Ki of 110 nM, thus still demonstrating a good fit of the amino acid residues into the protease binding pockets and also the importance of the flexibility of P1-P1' linkage for proper binding. A new type of peptide bond mimetic, N-hydroxylamine -CH2-N(OH)-, has been synthesized. Binding of hydroxylamino inhibitor of HIV-1 protease is further improved with respect to reduced-bond inhibitor.  相似文献   

6.
In human myeloperoxidase the heme is covalently attached to the protein via two ester linkages between the carboxyl groups of Glu242 and Asp94 and modified methyl groups on pyrrole rings A and C of the heme as well as a sulfonium ion linkage between the sulfur atom of Met243 and the beta-carbon of the vinyl group on pyrrole ring A. In the present study, wild-type recombinant myeloperoxidase (recMPO) and the variant Glu242Gln were produced in Chinese hamster ovary cells and investigated in a comparative sequential-mixing stopped-flow study in order to elucidate the role of the Glu242-heme ester linkage in the individual reaction steps of both the halogenation and peroxidase cycle. Disruption of the ester bond increased heme flexibility, blue shifted the UV-vis spectrum, and, compared with recMPO, decelerated cyanide binding (1.25 x 10(4) versus 1.6 x 10(6) M(-)(1) s(-)(1) at pH 7 and 25 degrees C) as well as compound I formation mediated by either hydrogen peroxide (7.8 x 10(5) versus 1.9 x 10(7) M(-)(1) s(-)(1)) or hypochlorous acid (7.5 x 10(5) versus 2.3 x 10(7) M(-)(1) s(-)(1)). The overall chlorination and bromination activity of Glu242Gln was 2.0% and 24% of recMPO. The apparent bimolecular rate constants of compound I reduction by chloride (65 M(-)(1) s(-)(1)), bromide (5.4 x 10(4) M(-)(1) s(-)(1)), iodide (6.4 x 10(5) M(-)(1) s(-)(1)), and thiocyanate (2.2 x10(5) M(-)(1) s(-)(1)) were 500, 25, 21, and 63 times decreased compared with recMPO. By contrast, Glu242Gln compound I reduction by tyrosine was only 5.4 times decreased, whereas tyrosine-mediated compound II reduction was 60 times slower compared with recMPO. The effects of exchange of Glu242 on electron transfer reactions are discussed.  相似文献   

7.
Geometry optimization and energy calculations have been performed at the density functional B3LYP/LANL2DZ level on hydrogen sulfide (HS-), dihydrogensulfide (H2S), thiomethanolate (CH3S-), thiomethanol (CH3SH), thiophenolate (C6H5S-), methoxyde (CH3O-), methanol (CH3OH), formiate (HCOO-), acetate (CH3COO-), carbonate (CO3(2-)), hydrogen carbonate (HCO3-), iminomethane (NH=CH2), [ZnS], [ZnS2]2-, [Zn(HS)]+, [Zn(H2S)]2+, [Zn(HS)4]2-, [Zn(CH3S)]+, [Zn(CH3S)2], [Zn(CH3S)3]-, [Zn(CH3S)4]2-, [Zn(CH3SH)]2+, [Zn(CH3SCH3)]2+, [Zn(C6H5S)]+, [Zn(C6H5S)2], [Zn(C6H5S)3]-, [Zn(HS)(NH=CH2)2]+, [Zn(HS)2(NH=CH2)2], [Zn(HS)(H2O)]+, [Zn(HS)(HCOO)], [Zn(HS)2(HCOO)]-, [Zn(CH3O)]+, [Zn(CH3O)2], [Zn(CH3O)3]-, [Zn(CH3O)4]2, [Zn(CH3OH)]2+, [Zn(HCOO)]+, [Zn(CH3COO)]+, [Zn(CH3COO)2], [Zn(CH3COO)3]-, [Zn(CO3)], [Zn(HCO3)]+, and [Zn(HCO3)(Imz)]+ (Imz, 1,3-imidazole). The computed Zn-S bond distances are 2.174A for [ZnS], 2.274 for [Zn(HS)]+, 2.283 for [Zn(CH3S)]+, and 2.271 for [Zn(C6H5S)]+, showing that sulfide anion forms stronger bonds than substituted sulfides. The nature of the substituents on sulfur influences only slightly the Zn-S distance. The optimized tetra-coordinate [Zn(HS)2(NH=CH2)2] molecules has computed Zn-S and Zn-N bond distances of 2.392 and 2.154A which compare well with the experimental values at the solid state obtained via X-ray diffraction for a number of complex molecules. The computed Zn-O bond distances for chelating carboxylate derivatives like [Zn(HOCOO)]+ (1.998A), [Zn(HCOO)]+ (2.021), and [Zn(CH3COO)]+ (2.001) shows that the strength of the bond is not much influenced by the substituent on carboxylic carbon atom and that CH3- and HO- groups have very similar effects. The DFT analysis shows also that the carboxylate Ligand has a preference for the bidentate mode instead of the monodentate one, at least when the coordination number is small.  相似文献   

8.
Payne JC  Rous BW  Tenderholt AL  Godwin HA 《Biochemistry》2003,42(48):14214-14224
Zinc binding to the two Cys(4) sites present in the DNA-binding domain (DBD) of nuclear hormone receptor proteins is required for proper folding of the domain and for protein activity. By utilizing Co(2+) as a spectroscopic probe, we have characterized the metal-binding properties of the two Cys(4) structural zinc-binding sites found in the DBD of human estrogen receptor alpha (hERalpha-DBD) and rat glucocorticoid receptor (GR-DBD). The binding affinity of Co(2+) to the two proteins was determined relative to the binding affinity of Co(2+) to the zinc finger consensus peptide, CP-1. Using the known dissociation constant of Co(2+) from CP-1, the dissociation constants of cobalt from hERalpha-DBD were calculated: K(d1)(Co) = 2.2 (+/- 1.0) x 10(-7) M and K(d2)(Co) = 6.1 (+/- 1.5) x 10(-7) M. Similarly, the dissociation constants of Co(2+) from GR-DBD were calculated: K(d1)(Co) = 4.1 (+/- 0.6) x 10(-7) M and K(d2)(Co) = 1.7 (+/- 0.3) x 10(-7) M. Metal-binding studies conducted in which Zn(2+) displaces Co(2+) from the metal-binding sites of hERalpha-DBD and GR-DBD indicate that Zn(2+) binds to each of the Cys(4) metal-binding sites approximately 3 orders of magnitude more tightly than Co(2+) does: the stoichiometric dissociation constants are K(d1)(Zn) = 1 (+/- 1) x 10(-10) M and K(d2)(Zn) = 5 (+/- 1) x 10(-10) M for hERalpha-DBD and K(d1)(Zn) = 2 (+/- 1) x 10(-10) M and K(d2)(Zn) = 3 (+/- 1) x 10(-10) M for GR-DBD. These affinities are comparable to those observed for most other naturally occurring structural zinc-binding sites. In contrast to the recent prediction by Low et. al. that zinc binding in these systems should be cooperative [Low, L. Y., Hernández, H., Robinson, C. V., O'Brien, R., Grossmann, J. G., Ladbury, J. E., and Luisi, B. (2002) J. Mol. Biol. 319, 87-106], these data suggest that the zincs that bind to the two sites in the DBDs of hERalpha-DBD and GR-DBD do not interact.  相似文献   

9.
B Imperiali  R H Abeles 《Biochemistry》1987,26(14):4474-4477
We have synthesized inhibitors of chymotrypsin, based on fluoromethyl ketones, that bind at S and S' subsites. "Small" inhibitors of serine proteases, which have previously been synthesized, only interact with S subsites. The parent compound is Ac-Leu-ambo-Phe-CF2H (1) (Ki = 25 X 10(-6) M). This inhibitor was modified by successively replacing H of the -CF2H group by -CH2CH2CONHCH3, (4), -CH2CH2CONH-Leu-NHMe (5), -CH2CH2CONH-Leu-Val-OEt (6), and -CH2CH2CONH-Leu-Arg-OMe (7). Corresponding Ki values are 7.8 (4), 0.23 (5), 0.21 (6), and 0.014 (7) microM. Extending 5 to 6 by addition of Val-OEt at P3' does not decrease Ki. In contrast, extension of 5 to 7 by incorporating Arg-OMe at P3' decreases Ki approximately 15-fold, suggesting interaction between Arg and the S3' subsite but no corresponding interaction at that subsite with Val. These results are in accordance with results obtained with the homologous family of avian ovomucoid third domain proteins. Proteins with Arg at the P3' position show highly favorable interactions with the protease at the S3' subsite [Park, S. J. (1985) Ph.D. Thesis, Purdue University; M. Laskowski, Jr., personal communication]. These results establish that incorporation of residues which interact with S' subsites significantly increases the efficacy of inhibitors and that valuable information concerning the most effective amino acid composition of small inhibitors can be obtained from the amino acid sequence of protein inhibitors.  相似文献   

10.
The antitumor platinum(II) compound, [Pt(dach)(Glu)] (dach=trans(+/-)-1,2-diaminocyclohexane, Glu=glutamate) was formulated with a stealth liposome to improve its biological activity. Liposomes were composed of PC/PEG2000-PE/CH (PC=1,2-diacyl-glycero-3-phosphocholine; PEG2000-PE=poly(ethylene glycol)2000-1,2-diacyl-glycero-3-phosphoethanolamine; CH=cholesterol) involving different acyl moieties of phospholipids such as DO (dioleoyl), DM (dimyristoyl) or DS (distearoyl) group. Among the different acyl groups in the stealth liposomes, the DM formulation was optimal for the preparation of the liposomal [Pt(dach)(Glu)] at the mole ratio of DMPC/PEG2000-DMPE/CH=50/5/45 and at the weight ratio of drug/lipid=1/20, which is represented as L-[Pt(dach)(Glu)]. In vitro cytotoxicity was examined in sensitive A2780 and ME180 and their cisplatin-resistant A2780/PDD and ME180/PDD cancer cells. L-[Pt(dach)(Glu)] was 2 approximately 3 times more cytotoxic than the free complex [Pt(dach)(Glu)] and cisplatin in sensitive cells, and 4 approximately 8 times more cytotoxic in resistant cells. Thus, the resistance index of L-[Pt(dach)(Glu)] was 1.3 approximately 2 while those of the free complex and cisplatin were 5 approximately 6, which indicates that L-[Pt(dach)(Glu)] overcome the cisplatin resistance in both resistant cells. In vivo antitumor activity was assayed against the L1210/S leukemia. The optimal activities (% T/C) of the free complex and L-[Pt(dach)(Glu)] were >459/20 and >442/200 mg/kg, respectively. Considering the amount of the platinum complex in L-[Pt(dach)(Glu)], the liposomal [Pt(dach)(Glu)] displayed 2-fold higher drug potency than the free complex. The biodistribution experiment using LE52 tumor-bearing mouse showed excellent lung targeting property of L-[Pt(dach)(Glu)].  相似文献   

11.
The cytotoxic lymphocyte serine proteinase granzyme B induces apoptosis of abnormal cells by cleaving intracellular proteins at sites similar to those cleaved by caspases. Understanding the substrate specificity of granzyme B will help to identify natural targets and develop better inhibitors or substrates. Here we have used the interaction of human granzyme B with a cognate serpin, proteinase inhibitor 9 (PI-9), to examine its substrate sequence requirements. Cleavage and sequencing experiments demonstrated that Glu(340) is the P1 residue in the PI-9 RCL, consistent with the preference of granzyme B for acidic P1 residues. Ala-scanning mutagenesis demonstrated that the P4-P4' region of the PI-9 RCL is important for interaction with granzyme B, and that the P4' residue (Glu(344)) is required for efficient serpin-proteinase binding. Peptide substrates based on the P4-P4' PI-9 RCL sequence and containing either P1 Glu or P1 Asp were cleaved by granzyme B (k(cat)/K(m) 9.5 x 10(3) and 1.2 x 10(5) s(-1) M(-1), respectively) but were not recognized by caspases. A substrate containing P1 Asp but lacking P4' Glu was cleaved less efficiently (k(cat)/K(m) 5.3 x 10(4) s(-1) M(-1)). An idealized substrate comprising the previously described optimal P4-P1 sequence (Ile-Glu-Pro-Asp) fused to the PI-9 P1'-P4' sequence was efficiently cleaved by granzyme B (k(cat)/K(m) 7.5 x 10(5) s(-1) M(-1)) and was also recognized by caspases. This contrasts with the literature value for a tetrapeptide comprising the same P4-P1 sequence (k(cat)/K(m) 6.7 x 10(4) s(-1) M(-1)) and confirms that P' residues promote efficient interaction of granzyme B with substrates. Finally, molecular modeling predicted that PI-9 Glu(344) forms a salt bridge with Lys(27) of granzyme B, and we showed that a K27A mutant of granzyme B binds less efficiently to PI-9 and to substrates containing a P4' Glu. We conclude that granzyme B requires an extended substrate sequence for specific and efficient binding and propose that an acidic P4' substrate residue allows discrimination between early (high affinity) and late (lower affinity) targets during the induction of apoptosis.  相似文献   

12.
Liu ZX  Fei H  Chi CW 《FEBS letters》2004,556(1-3):116-120
Eglin c with mutants L45R and D42R at the P(1) and P(4) positions has been reported to become a stable inhibitor toward the proprotein convertases (PC), furin and kexin, with a K(i) of 2.3x10(-8) and 1.3x10(-10) M, respectively. The mutant was further engineered at the P(2)'-P(4)' positions to create a more potent and selective inhibitor for each enzyme. The residue Asp at P(1)' which is crucial for stabilizing the conformation of eglin c remained unchanged. The eglin c mutants cloned into the vector pGEX-2T and expressed in Escherichia coli (DH5alpha) were purified to homogeneity, and their inhibitory activities toward the purified recombinant furin and kexin were examined. The results showed that (1) Leu47 at P(2)' replaced with either a positively or negatively charged residue resulted in a decrease in inhibitory activities to both enzymes; (2) the replacement of Arg with Asp at P(3)' was favorable for inhibiting furin with a K(i) of 7.8 x 10(-9) M, but not for inhibiting kexin; (3) the replacement of Tyr with Glu at P(4)' increased the inhibitory activity to kexin with a K(i) of 3 x 10(-11) M, but was almost without any influence on furin inhibition. It was indicated that the inhibitory specificity of eglin c could be changed from inhibiting elastase to inhibiting PCs by site-directed mutation at the P positions, while the inhibitory selectivity to furin or kexin could be optimized by mutation at the P' positions.  相似文献   

13.
New synthetic mercaptotripeptides (HS-CH2-CH2-CO-Pro-Yaa) which inhibit Achromobacter iophagus collagenase were produced in order to obtain more powerful bacterial collagenase inhibitors than currently available, and to investigate the specificity of the S3' subsite of the enzyme. Since similar binding constants were found for inhibitors carrying uncharged residues of various sizes in the P3' position (Yaa = Ala, Leu, Phe, Pro, Hyp) steric hindrance at the collagenase S3' appears relatively limited. The compound (HS-CH2-CH2-CO-Pro-Arg), which carries an arginine residue in the position P3' and had the highest inhibition constant of the series tested (Ki = 0.5 microM), proved to be the strongest inhibitor so far reported in the literature. The weakest in the present series was the compound (HS-CH2-CH2-CO-Pro-Asp) which carries an aspartic residue in position P3' and had a Ki = 70 microM. The present work revealed that the charged groups in the P3' position play a key role in the interaction of the inhibitors with the enzyme.  相似文献   

14.
A decapeptide corresponding to the entire amino acid sequence of neurokinin A, a porcine spinal cord peptide, was synthesized in a conventional manner using protecting groups removable by 1 M TFMSA-thioanisole in TFA. The HS-CH2CH2CO group was introduced onto the synthetic neurokinin A by reaction of 3-(S-acetyl-thiopropionyl)-thiazolidine-2-thione, followed by deacetylation with hydroxylamine. 2,4-Dinitrophenyl-p-(beta-nitrovinyl)-benzoate trapped the above HS-CH2CH2CO-neurokinin A derivative in acidic media, then BSA in basic media in nearly quantitative yield. A similar decapeptide, neurokinin B, was also synthesized and conjugated onto BSA using an alternative SH-introducing reagent, 3-(S-p-methoxybenzyl-thiopropionyl)-thiazolidine-2-thione, and the above heterobifunctional conjugating reagent.  相似文献   

15.
A water-insoluble alpha-(1-->3)-D-glucan (A) from Lentinus edodes was fractionated into 13 fractions in dimethyl sulfoxide containing 0.25 M lithium chloride (0.25 M LiCl-Me(2)SO). Five fractions were treated with sulfur trioxide-pyridine complex at 25 degrees C to synthesize water-soluble sulfated derivatives (S-A). The weight-average molecular weights, M(w), and intrinsic viscosities [eta], of the samples A and S-A were determined by multi-angler laser light scattering (MALLS), and viscosity. The M(w) dependence of [eta] and of the radius of gyration (z)(1/2), was found to be represented approximately by [eta]=4.9 x 10(-2) M(w)(0.67) (cm(3) g(-1)), and (z)(1/2)=4.8 x 10(-2) M(w)(0.54) (nm) for the alpha-glucan in 0.25 M LiCl-Me(2)SO in the M(w) range from 7.24 x 10(4) to 4.21 x 10(5), and by [eta]=6.8 x 10(-4) M(w) 1.06 (cm(3) g(-1)), and (z)(1/2)=9.4 x 10(-4) M(w)(0.92) (nm) for the sulfated alpha-glucan in aqueous 0.5 M NaCl in the M(w) range from 5.92 x 10(4) to 1.42 x 10(5) at 25 degrees C. The results indicate that the alpha-(1-->3)-D-glucan exists as a flexible chain in 0.25 M LiCl-Me(2)SO, and its sulfated derivative in 0.5 M NaCl aqueous has stiffer chains than the original. (13)C NMR indicated that intramolecular hydrogen bonding occurred in the sulfated alpha-glucan, causing the observed chain stiffness.  相似文献   

16.
Shih I  Been MD 《Biochemistry》2000,39(31):9055-9066
A minimal kinetic mechanism for a trans-acting ribozyme derived from the HDV antigenomic RNA self-cleaving element was established from steady-state, pre-steady-state, single-turnover, and binding kinetics. Rate constants for individual steps, including substrate binding and dissociation, cleavage, and product release and binding, were measured at 37 degrees C at pH 8.0 in 10 mM Mg(2+) using oligonucleotides as either substrates, noncleavable analogues or 3' product mimics. A substrate containing a normal 3',5'-linkage was cleaved with a first-order rate constant (k(2)) of 0.91 min(-)(1). The association rate constant for the substrate to the ribozyme (2.1 x 10(7) M(-)(1) min(-)(1)) was at the lower range of the expected value for RNA duplex formation, and the substrate dissociated with a rate constant (1.4 min(-)(1)) slightly faster than that for cleavage. Thus the binary complex was not at equilibrium with free enzyme and substrate prior to the cleavage step. Following cleavage, product release was kinetically ordered in that the 5' product was released rapidly (>12 min(-)(1)) relative to the 3' product (6.0 x 10(-)(3) min(-)(1)). Rapid 5' product release and lack of a demonstrable binding site for the 5' product could contribute to the difficulty in establishing the ribozyme-catalyzed reverse reaction (ligation). Slow release of the 3' product was consistent with the extremely low turnover under steady-state conditions as 3' product dissociation was rate-limiting. The equilibrium dissociation constant for the substrate was 24-fold higher than that of the 3' cleavage product. A substrate with a 2',5'-linkage at the cleavage site was cleaved with a rate constant (k(2)) of 1.1 x 10(-)(2) min(-)(1). Thus, whereas cleavage of a 3',5'-linkage followed a Briggs-Haldane mechanism, 2', 5' cleavage followed a Michaelis-Menten mechanism.  相似文献   

17.
We present measurements of low-energy (<18 eV) electron-stimulated desorption of anions from acetamide (CH(3)CONH(2)) and dimethyl disulfide [DMDS: (CH(3)S)(2)] films. Electron irradiation of physisorbed CH(3)CONH(2) produces H(-), CH(3)(-) and O(-) anions, whereas the H(-), CH(2)(-), CH(3)(-), S(-), SH(-) and SCH(3)(-) anions are observed to desorb from the DMDS film. Below 12 eV, the dependence of the anion yields on the incident electron energy exhibits structures that indicate that a resonant process (i.e. dissociative electron attachment) is responsible for molecular fragmentation. Within the range of 1-18 eV, it is found that (1.7 and 1.4) x 10(7) H(-) ions/incident electron and (7.8 x 10(-11) and 4.3 x 10(-8)) of the other ions/incident electron are desorbed from acetamide and DMDS films, respectively. These results suggest that, within proteins, the disulfide bond is more sensitive to low-energy electron attack than the peptide bond. In biological cells, some proteins interact closely with nucleic acid. Therefore, the observed fragments, when produced from secondary low-energy electrons generated by high-energy radiation, not only may denature proteins, but may also induce reactions with the nearby nucleic acid and damage DNA.  相似文献   

18.
Matsuno H  Furusawa H  Okahata Y 《Biochemistry》2005,44(7):2262-2270
Catalytic DNA cleavage reactions by an ATP-dependent deoxyribonuclease (DNase) from Micrococcus luteus were monitored directly with a DNA-immobilized 27-MHz quartz-crystal microbalance (QCM). The 27-MHz QCM is a very sensitive mass-measuring device in aqueous solution, as the frequency decreases linearly with increasing mass on the electrode at a nanogram level. Three steps in ATP-dependent DNA hydrolysis reactions, including (1) binding of DNase to the end of double-stranded DNA (dsDNA) on the QCM electrode (mass increase), (2) degradation of one strand of dsDNA in the 3' --> 5' direction depending on ATP (mass decrease), and (3) release of the enzyme from the nonhydrolyzed 5'-free-ssDNA (mass decrease), could be monitored stepwise from the time dependencies of QCM frequency changes. Kinetic parameters for each step were obtained as follows. The binding constant (K(a)) of DNase to the dsDNA was determined as (28 +/- 2) x 10(6) M(-)(1) (k(on) = (8.0 +/- 0.3) x 10(3) M (-)(1) s(-)(1) and k(off) = (0.29 +/-0.01) x 10(-)(3) s(-)(1)), and it decreased to (0.79 +/- 0.16) x 10(6) M(-)(1) (k'(on) = (2.3 +/- 0.2) x 10(3) M (-)(1) s(-)(1) and k'(off) = (2.9 +/- 0.1) x 10(-)(3) s(-)(1)) for the completely nonhydrolyzed 5'-free ssDNA. This is the reason the DNase bound to the dsDNA substrate can easily release from the nonhydrolyzed 5'-free-ssDNA after the complete hydrolysis of the 3' --> 5' direction of the complementary ssDNA. K(a) values depended on the DNA structures on the QCM, and the order of these values was as follows: the dsDNA having a 4-base-mismatched base-pair end (3) > the dsDNA having a 5' 15-base overhanging end (2) > the dsDNA having a blunt end (1) > the ssDNA having a 3'-free end (4) > the ssDNA having a 5'-free end (5). Thus, DNase hardly recognized the free 5' end of ssDNA. Michaelis-Menten parameters (K(m) for ATP and k(cat)) of the hydrolysis process also could be obtained, and the order of k(cat)/K(m) was as follows: the dsDNA having a blunt end (1) approximately the dsDNA having a 4-base-mismatched base-pair end (3) > the ssDNA having a free 3' end (4) > the ssDNA having a free 5' end (5). Thus, DNase could not recognize and not hydrolyze the free 5' end of ssDNA. The DNA hydrolysis reaction could be driven by dATP and GTP (purine base) as well as ATP, whereas the cleavage efficiency was very low driven with UTP, CTP (pyrimidine base), ADP, and AMP.  相似文献   

19.
20.
C A Hastings  J K Barton 《Biochemistry》1999,38(31):10042-10051
Metallointercalator-peptide conjugates that provide small molecular mimics to explore peptide-nucleic acid recognition have been prepared. Specifically, a family of peptide conjugates of [Rh(phi)(2)(phen')](3+) [where phi = 9,10-phenanthrenequinone diimine and phen' = 5-(amidoglutaryl)-1,10-phenanthroline] has been synthesized and their DNA-binding characteristics examined. Single amino acid modifications were made from the parent metallointercalator-peptide conjugate [Rh(phi)(2)(phen')](3+)-AANVAIAAWERAA-CONH(2), which targets 5'-CCA-3' site-specifically. Moving the glutamate at position 10 in the sequence of the appended peptide to position 6 {[Rh(phi)(2)(phen')](3+)-AANVAEAAWARAA-CONH(2)} changed the sequence preference of the metallointercalator-peptide conjugate to 5'-ACA-3'. Subsequent mutation of the glutamate at position 6 to arginine {[Rh(phi)(2)(phen')](3+)-AANVARAAWARAA-CONH(2)} caused more complex changes in DNA recognition. Thermodynamic dissociation constants were determined for these metallointercalator-peptide conjugates by photoactivated DNA cleavage assays with the rhodium intercalators. At 55 degrees C in the presence of 5 mM MnCl(2), [Rh(phi)(2)(phen')](3+)-AANVAIAAWERAA-CONH(2) binds to a 5'-CCA-3' site with K(d) = 5.7 x 10(-)(8) M, whereas [Rh(phi)(2)(phen')](3+)-AANVAEAAWARAA-CONH(2) binds to its target 5'-ACA-3' site with K(d) = 9.9 x 10(-8) M. The dissociation constant for [Rh(phi)(2)(phen')](3+) with random-sequence DNA is 7.0 x 10(-7) M. Structural models have been developed and refined to account for the observed sequence specificities. As with much larger DNA-binding proteins, with these metal-peptide conjugate mimics, single amino acid changes can lead to single or multiple base changes in the DNA site targeted.  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