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 15 毫秒
1.
The participation of a low-barrier hydrogen bond (LBHB) in the mechanism of action of chymotrypsin introduces a new role for Asp 102 and His 57 in catalysis [C. S. Cassidy, J. Lin, and P. A. Frey (1997) Biochemistry 36, 4576-4584]. It is postulated that the LBHB increases the basicity of His 57-N(epsilon2) in the transition state, thereby facilitating the abstraction of a proton from Ser 195, and stabilizes the tetrahedral intermediate in the acylation step. Evidence for this mechanism includes the downfield chemical shift of the proton bridging His 57 and Asp 102 in transition-state analog complexes and the low deuterium fractionation factors for this proton in the same complexes. We present additional spectroscopic evidence supporting the assignment of an LBHB between His 57 and Asp 102. The tetrahedral addition complex between Ser 195 of chymotrypsin and N-acetyl-l-leucyl-l-phenylalanyl trifluoromethylketone is regarded as a close structural analog of a tetrahedral intermediate. The deuterium NMR signal for the downfield deuteron bridging His 57 and Asp 102 in D(2)O has now been observed as a broad band centered at 17.8 +/- 0.5 ppm. The proton NMR signal in H(2)O is centered at 18.9 +/- 0.05 ppm. The two signals are clearly separated corresponding to a deuterium isotope effect of Delta[delta(H) - delta(D)] = 1.1 +/- 0.5 ppm. Deuterium isotope effects in this range are characteristic of LBHBs, and this observation provides further support for the assignment of the proton bridging His 57 and Asp 102 in transition-state analog complexes as an LBHB.  相似文献   

2.
Ishida T 《Biochemistry》2006,45(17):5413-5420
To elucidate the catalytic advantage of the low-barrier hydrogen bond (LBHB), we analyze the hydrogen bonding network of the catalytic triad (His57-Asp102-Ser195) of serine protease trypsin, one of the best examples of the LBHB reaction mechanism. Especially, we focus on the correlation between the change of the chemical shifts and the structural rearrangement of the active site in the acylation process. To clarify LBHB, we evaluate the two complementary properties. First, we calculate the NMR chemical shifts of the imidazole ring of His57 by the gauge-including atomic orbital (GIAO) approach within the ab initio QM/MM framework. Second, the free energy profile of the proton transfer from His57 to Asp102 in the tetrahedral intermediate is obtained by ab initio QM/MM calculations combined with molecular dynamics free energy perturbation (MD-FEP) simulations. The present analyses reveal that the calculated shifts reasonably reproduce the observed values for (1)H chemical shift of H(epsilon)(1) and H(delta)(1) in His57. The (15)N and (13)C chemical shifts are also consistent with the experiments. It is also shown that the proton between His57 and Asp102 is localized at the His57 side. This largely downfield chemical shift is originated from the strong electrostatic interaction, not a covalent-like bonding character between His57 and Asp102. Also, it is proved that a slight downfield character of H(epsilon)(1) is originated from a electrostatic interaction between His57 and the backbone carbonyl group of Val213 and Ser214. These downfield chemical shifts are observed only when the tetrahedral intermediate is formed in the acylation process.  相似文献   

3.
K Brady  A Z Wei  D Ringe  R H Abeles 《Biochemistry》1990,29(33):7600-7607
The peptidyl trifluoromethyl ketones Ac-Phe-CF3 (1) and Ac-Leu-Phe-CF3 (2) are inhibitors of chymotrypsin. They differ in Ki (20 and 2 microM, respectively) as well as in their kinetics of association with chymotrypsin in that 1 is rapidly equilibrating, with an association rate too fast to be observed by steady-state techniques, while 2 is "slow binding", as defined by Morrison and Walsh [Morrison, J. F., & Walsh, C. T. (1988) Adv. Enzymol. Relat. Areas Mol. Biol. 61, 202], with a second-order association rate constant of 750 M-1 s-1 at pH 7.0 [Imperiali, B., & Abeles, R. (1986) Biochemistry 25, 3760]. The crystallographic structures of the complexes of gamma-chymotrypsin with inhibitors 1 and 2 have been determined in order to establish whether structural or conformational differences can be found which account for different kinetic and thermodynamic properties of the two inhibitors. In both complexes, the active-site Ser 195 hydroxyl forms a covalent hemiketal adduct with the trifluoromethyl ketone moiety of the inhibitor. In both complexes, the trifluoromethyl group is partially immobilized, but differences are observed in the degree of interaction of fluorine atoms with the active-site His 57 imidazole ring, with amide nitrogen NH 193, and with other portions of the inhibitor molecule. The enhanced potency of Ac-Leu-Phe-CF3 relative to Ac-Phe-CF3 is accounted for by van der Waals interactions of the leucine side chain of the inhibitor with His 57 and Ile 99 side chains and by a hydrogen bond of the acetyl terminus with amide NH 216 of the enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The naturally occurring peptidyl protease inhibitor leupeptin (N-acetyl-L-leucyl-L-leucyl-L-argininal) has been prepared labeled with 13C at the argininal carbonyl. 13C chemical shift data for the trypsin-leupeptin inhibitor complex in the pH range 3.0-7.6 reveal the presence of two pH-dependent covalent complexes, suggestive of two interconverting diastereomers at the new asymmetric tetrahedral center created by covalent addition of Ser195 to either side of the 13C-enriched aldehyde of the inhibitor. At pH 7 two signals are observable at delta 98.8 and delta 97.2 (84:16 ratio), while at pH 3.0 the latter signal predominates. In the selective proton 13C-edited NOE spectrum of the major diastereomer at pH 7.4, a strong NOE is observed between the hemiacetal proton of the inhibitor and the C2 proton of His57 of the enzyme, thus defining the stereochemistry of the high pH complex to the S configuration in which the hemiacetal oxygen resides in the oxyanion hole. pH titration studies further indicate that the 13C chemical shift of the S diastereomer follows a titration curve with a pKa of 4.69, the magnitude of which is consistent with direct titration of the hemiacetal oxygen. Similar pH-dependent chemical shifts were obtained by using CPMAS 13C NMR, providing evidence for the existence of the same diastereomeric equilibrium in the solid state.  相似文献   

5.
K Brady  T C Liang  R H Abeles 《Biochemistry》1989,28(23):9066-9070
The effects of pH on the kinetics of association and dissociation of chymotrypsin and the dipeptidyl trifluoromethyl ketone (TFK) N-acetyl-L-leucyl-L-phenylalanyltrifluoromethane (1) were examined through the pH range 4-9.5. The pH dependence of the association rate (kon) is similar to that of kcat/Km for ester and peptide substrates and is dependent on two pK's at 7.0 and 8.9. We assign these pK's to the active site His and to the amino group of the N-terminal isoleucine residue. Ki for the complex of 1 and chymotrypsin has a pH dependence very similar to that of kon, and we conclude that the same ionizable groups which determine the pH dependence of kon are involved. The dissociation constant of the enzyme-inhibitor complex (koff) shows no pH dependence between pH 4 and pH 9.5. The data indicate that the inhibitor reacts with a form of the enzyme in which His 57 is unprotonated, and the resulting complex contains no groups which ionize between pH 4 and pH 9.5. This is consistent with conclusions previously reached from NMR data (Liang & Abeles, 1987). These experiments led to the conclusion that 1 reacts with chymotrypsin to form a tetrahedral complex in which His 57 is protonated (pK greater than 9.5) and the OH group of serine 195 has added to the carbonyl group of 1 to form an ionized hemiketal (pK less than 4.9). The pK of His 57 is increased by greater than 3 units over that in the free enzyme, and the pK of the hemiketal decreased by greater than 4 units compared to the pK in solution.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Generation of diastereomeric phosphonate ester adducts of chymotrypsin was evidenced for the first time by 31P NMR and spectrophotometric kinetic measurements. 31P NMR signals were recorded for 4-nitrophenyl 2-propyl methylphosphonate (IMN) at 32.2 ppm and for its hydrolysis product at 26.3 ppm downfield from phosphoric acid. The inhibition of α-chymotrypsin at pH > 8.0 by the faster reacting enantiomer of IMN or 2-propyl methylphosphonochloridate (IMCl), or other phosphonate ester analogs of these compounds, all caused a ~6.0 ppm downfield shift of the 31P signal to the 39–40 ppm region. IMN, when applied below the stoichiometric amount of chymotrypsin, under the same conditions, generated two signals, at 39.0 and at 37.4 ppm. Scans accumulated in hourly intervals showed the decomposition of both diastereomers, with approximate half-lives of 12 h at pH 8.0 and 22°C, into a species with a resonance at 35.5 ppm. The most likely reaction to account for the appearance of this new peak is the enzymic dealkylation of the isopropyl group from the covalently bound phosphonate ester. We base this conclusion mostly on the similarity of the upfield shift to the hydrolysis of phosphonate esters. Contrary to experience with phosphate ester adducts of serine proteases, no signal was detected higher than 25.0 ppm downfield from phosphoric acid for several phosphonate ester adducts of chymotrypsin and in no case did the resonance for the adduct shift further downfield in the course of the experiments. © 1993 Wiley-Liss, Inc.  相似文献   

7.
T C Liang  R H Abeles 《Biochemistry》1987,26(24):7603-7608
A dipeptidyl trifluoromethyl ketone, N-acetyl-L-leucyl-L-[1-13C]phenylalanyl trifluoromethyl ketone, was synthesized. This compound inhibits chymotrypsin with Ki = 1.2 microM [Imperiali B., & Abeles, R.H. (1986) Biochemistry 25, 3760-3767]. The complex formed between this inhibitor and alpha-chymotrypsin was examined with 1H, 13C, and 19F NMR spectroscopy to establish its structure in solution. The keto group of the trifluoro ketone is present as an ionized hemiketal group as deduced from the comparison of its 13C chemical shift with those of model hemiketals. The pKa of the hemiketal hydroxyl in the complex is approximately 4.9, which is about 4.2 units lower than the pKa of model hemiketals. This observation provides direct evidence that serine proteases are able to stabilize the oxyanions of tetrahedral adducts. Evidence is also presented for the presence of an Asp-His H bond and protonation of the imidazole group of His-57 in the tetrahedral adduct. The pKa of His-57 is higher than 10. This observation directly indicates that the pKa of His-57 is elevated in a complex containing a tetrahedral adduct.  相似文献   

8.
Z-Ala-Pro-Phe-glyoxal (where Z is benzyloxycarbonyl) has been shown to be a competitive inhibitor of subtilisin with a K(i)=2.3+/-0.2 microM at pH 7.0 and 25 degrees C. Using Z-Ala-Pro-[2-(13)C]Phe-glyoxal we have detected a signal at 107.3 ppm by (13)C NMR, which we assign to the tetrahedral adduct formed between the hydroxy group of serine-195 and the (13)C-enriched keto-carbon of the inhibitor. The chemical shift of this signal is pH independent from pH 4.2 to 7.0 and we conclude that the oxyanion pK(a)<3. This is the first observation of oxyanion formation in a reversible subtilisin-inhibitor complex. The inhibitor is bound as a hemiketal which is in slow exchange with the free inhibitor. Inhibitor binding depends on a pK(a) of approximately 6.5 in the free enzyme and on a pK(a)<3.0 when the inhibitor is bound to subtilisin. Protonation of the oxyanion promotes the disassociation of the inhibitor. We show that oxyanion formation cannot be rate limiting during catalysis and that subtilisin stabilises the oxyanion by at least 45.1 kJ mol(-1). We conclude that if the energy required for oxyanion stabilisation is utilised as binding energy in drug design it should make a significant contribution to inhibitor potency.  相似文献   

9.
A range of substrate-derived chloromethane inhibitors have been synthesized which have one to four amino acid residues. These have been used to inhibit both subtilisin and chymotrypsin. Using 13C NMR, we have shown that all except one of these inhibitors forms a tetrahedral adduct with chymotrypsin, subtilisin, and trypsin. From the pH-dependent changes in the chemical shift of the hemiketal carbon of the tetrahedral adduct, we are able to determine the oxyanion pKa in the different inhibitor derivatives. Our results suggest that in both the subtilisin and chymotrypsin chloromethane derivatives the oxyanion pKa is largely determined by the type of amino acid residue occupying the S1, subsite while binding in the S2-S4 subsites only has minor effects on oxyanion pKa values. Using free energy relationships, we determine that the different R groups of the amino acid residues binding in the S1 subsite only have minor effects on the oxyanion pKa values. We propose that the lower polarity of the chymotrypsin active site relative to that of the subtilisin active site explains why the oxyanion pKa is higher and more sensitive to the type of chloromethane inhibitor used in the chymotrypsin derivatives than in the subtilisin derivatives.  相似文献   

10.
This work describes in-depth NMR characterization of a unique low-barrier hydrogen bond (LBHB) between an active site residue from the enzyme and a bound inhibitor: the complex between secreted phospholipase A(2) (sPLA(2), from bee venom and bovine pancreas) and a transition-state analog inhibitor HK32. A downfield proton NMR resonance, at 17-18 ppm, was observed in the complex but not in the free enzyme. On the basis of site-specific mutagenesis and specific 15N-decoupling, this downfield resonance was assigned to the active site H48, which is part of the catalytic dyad D99-H48. These results led to a hypothesis that the downfield resonance represents the proton (H(epsilon 2) of H48) involved in the H-bonding between D99 and H48, in analogy with serine proteases. However, this was shown not to be the case by use of the bovine enzyme labeled with specific [15N(epsilon 2)]His. Instead, the downfield resonance arises from H(delta1) of H48, which forms a hydrogen bond with a non-bridging phosphonate oxygen of the inhibitor. Further studies showed that this proton displays a fractionation factor of 0.62(+/-0.06), and an exchange rate protection factor of >100 at 285 K and >40 at 298 K, which are characteristic of a LBHB. The pK(a) of the imidazole ring of H48 was shown to be shifted from 5.7 for the free enzyme to an apparent value of 9.0 in the presence of the inhibitor. These properties are very similar to those of the Asp em leader His LBHBs in serine proteases. Possible structural bases and functional consequences for the different locations of the LBHB between these two types of enzymes are discussed. The results also underscore the importance of using specific isotope labeling, rather than extrapolation of NMR results from other enzyme systems, to assign the downfield proton resonance to a specific hydrogen bond. Although our studies did not permit the strength of the LBHB to be accurately measured, the data do not provide support for an unusually strong hydrogen bond strength (i.e. >10 kcal/mol).  相似文献   

11.
13C NMR is used to detect ionizations within a trypsin-chloromethyl ketone inhibitor complex. The pKa values observed are compared with those predicted by free-energy relationships. For the denatured/autolyzed inhibitor complex, a pKa = 5.26 is observed, which is assigned to the ionization of the imidazole of histidine-57. For the intact inhibitor complex a pKa = 7.88 is determined. This pKa is assigned to the ionization of the hemiketal hydroxyl (pKa = 7.88-8.1) and provides the first direct evidence that the serine proteases are able to stabilize the oxyanion of tetrahedral adducts. Indirect evidence is adduced that the imidazole pK1 of histidine-57 is greater than or equal to 8.1. Line-broadening studies suggest that there may be extra fast exchange line broadening, which could result from rapid tautomeric exchange between neutral and zwitterionic species within the inhibitor complex. The significance of these results for the catalytic mechanism of serine proteases is discussed.  相似文献   

12.
Carbon-13 (13C) nuclear magnetic resonance spectroscopy (NMR) is performed to characterize the formation of carbamino adducts between insulin and (13C) carbon dioxide over a range of pH values in the presence of a physiological concentration (23 mM) of sodium bicarbonate. The peaks from two of the carbamino adducts resonate at higher frequencies than the signal from bicarbonate, at 164.6 and 165.3 ppm, and are attributed to the adducts with the terminal amino groups of phenylalanine B1 and glycine A1. The intensities of these signals vary with the pH, with unique patterns. Over 6% of each terminal amino group exists as the carbamino adduct at the optimum pH values of 7.8 and 8.3. A unique third adduct resonates at 159.3 ppm, and is attributed to lysine B29. This adduct is present on 2% of the insulin molecules at pH 8.2, but has minimal intensity at pH 7.4. No signals from adducts are detected below pH 6.2, where the amino groups exist predominantly in the protonated form. Creation of the adducts is rapid and they are stable for over 4 wk at 37 degrees C. The narrow bandwidth of the resonance of the adduct (4.0-4.5 Hz) relative to the irreversible cyanate adduct is consistent with molecular forms of the carbamino adduct smaller than the 2-Zn-hexamer which is the preponderate form of clinically utilized U-100 insulin (i.e., 100 U/ml).  相似文献   

13.
K Brady  R H Abeles 《Biochemistry》1990,29(33):7608-7617
A series of seven peptidyl trifluoromethyl ketone (TFK) inhibitors of chymotrypsin have been prepared which differ at the P1 and P2 subsites. Inhibition equilibria and kinetics of association and dissociation with chymotrypsin have been measured. The association rate of Ac-Phe-CF3 was measured at enzyme concentrations between 8 nM and 117 microM in order to examine the relation between the ketone/hydrate equilibrium of trifluoromethyl ketones and the "slow binding" by these inhibitors. The association rate decreases at high enzyme concentrations, indicating that TFK ketone is the reactive species and that conversion of TFK hydrate to ketone becomes rate limiting under these conditions. Inhibitors with hydrophobic side chains at P2 bind more tightly but more slowly to chymotrypsin, indicating that formation of van der Waals contacts between the P2 side chain and the His 57 and Ile 99 side chains of chymotrypsin is a relatively slow process. Inhibitor properties were compared to the Michaelis-Menten kinetic constants of a homologous series of peptide methyl ester and peptide amide substrates. Plots of log Ki vs log (kcat/Km) are linear with slopes of 0.65 +/- 0.2, indicating that these inhibitors are able to utilize 65% of the total binding energy between chymotrypsin and its hydrolytic transition state.  相似文献   

14.
The proton nuclear magnetic resonance signal of the His57-Asp102 hydrogen bonded proton in the charge relay system of chymotrypsinogen A and chymotrypsin Aδ has been monitored to determine the influence of substrate analogues and competitive inhibitors on the electronic state of the active site regions. Borate ion, benzene boronic acid and 2-phenylethylboronic acid, when bound to chymotrypsin at pH 9.5 shift the resonance position of the His-Asp hydrogen bonded proton to ?15.9, ?16.3 and ?17.2 parts per million, respectively. These positions are intermediate between the low pH position in the free enzyme of ?18.0 parts per million and the high pH position of ?14.9 parts per million. The presence of these analogues prevents the His-Asp proton resonance from titrating in the region of pH 6 to 9.5. Similar low field shifts are observed for the hydrogen bonded proton resonance of subtilisin BPN′ when complexed with these boronic acids. The results support the chemical and crystallographic data which show that negatively charged tetrahedral adducts of the boronic acid substrate analogues are formed at the active sites of these enzymes. When combined with similar nuclear magnetic resonance data for the binding of N-acetyl-l-tryptophan to chymotrypsin Aδ, they suggest that a direct interaction occurs between the active site histidine and the atom occupying the leaving group position of the substrate, presumably a hydrogen bond.The His-Asp proton resonance was also monitored in complexes of chymotrypsin Aδ with bovine pancreatic trypsin inhibitor over the pH range 4 to 9. In the complex the low field proton resonance had a field position of ?14.9 parts per million over the pH range 4 to 9 indicating that His57 is in the neutral form, similar to the active enzyme at high pH.  相似文献   

15.
Z-Ala-Pro-Phe-glyoxal (where Z is benzyloxycarbonyl) has been shown to be a competitive inhibitor of subtilisin with a Ki=2.3±0.2 μM at pH 7.0 and 25 °C. Using Z-Ala-Pro-[2-13C]Phe-glyoxal we have detected a signal at 107.3 ppm by 13C NMR, which we assign to the tetrahedral adduct formed between the hydroxy group of serine-195 and the 13C-enriched keto-carbon of the inhibitor. The chemical shift of this signal is pH independent from pH 4.2 to 7.0 and we conclude that the oxyanion pKa<3. This is the first observation of oxyanion formation in a reversible subtilisin–inhibitor complex. The inhibitor is bound as a hemiketal which is in slow exchange with the free inhibitor. Inhibitor binding depends on a pKa of ~6.5 in the free enzyme and on a pKa<3.0 when the inhibitor is bound to subtilisin. Protonation of the oxyanion promotes the disassociation of the inhibitor. We show that oxyanion formation cannot be rate limiting during catalysis and that subtilisin stabilises the oxyanion by at least 45.1 kJ mol?1. We conclude that if the energy required for oxyanion stabilisation is utilised as binding energy in drug design it should make a significant contribution to inhibitor potency.  相似文献   

16.
Propylene oxide (PO) is a direct-acting mutagen and rodent carcinogen. We have studied how PO modifies 2'-deoxynucleosides at pH 7.0-7.5 and 37 degrees C for 10 h. PO reacts as an SN2 alkylating agent by forming the following 2-hydroxypropyl (HP) adducts: N6-HP-dAdo (7% yield), 7-HP-Gua (37%) and 3-HP-dThd (4%). Alkylation at N-3 of dCyd resulted in conversion of the adjacent exocyclic imino group at C-4 to an oxygen (hydrolytic deamination) with the formation of a dUrd adduct, 3-HP-dUrd (14%). Ultraviolet spectroscopy and mass spectrometry were used for the structural determination of these adducts. Confirmation of the unexpected 3-HP-dUrd adduct was provided by an accurate mass measurement technique where diagnostic ions in the mass spectra of 3-HP-dUrd were measured to within 0.0005 atomic mass units of the predicted mass. PO was reacted in vitro with calf thymus DNA (pH 7.0-7.5, 37 degrees C, 10 h) and yielded N6-HP-dAdo (1 nmol/mg DNA), 3-HP-Ade (14 nmol/mg DNA), 7-HP-Gua (133 nmol/mg DNA) and 3-HP-dUrd (13 nmol/mg DNA). A mechanism for the hydrolytic deamination of 3-HP-dCyd to 3-HP-dUrd involving the OH on the HP side chain is proposed. This cytosine to uracil conversion may play a role in the mutagenic and carcinogenic activity of this epoxide.  相似文献   

17.
P I Bauer  A Hakam  E Kun 《FEBS letters》1986,195(1-2):331-338
Calf thymus and rat liver poly(ADP-ribose) polymerase enzymes, and the polymerase present in extracts of rat liver nuclei synthesize unstable mono-ADP-ribose protein adducts at 100 nM or lower NAD concentrations. The isolated enzyme-mono-ADP-ribose adduct hydrolyses to ADP-ribose and enzyme protein at pH values slightly above 7.0 indicating a continuous release of ADP-ribose from NAD through this enzyme-bound intermediate under physiological conditions. NH2OH at pH 7.0 hydrolyses the mono-ADP-ribose enzyme adduct. Desamino NAD and some other homologs at nanomolar concentrations act as 'forward' activators of the initiating mono-ADP-ribosylation reaction. These NAD analogs at micromolar concentrations do not affect polymer formation that takes place at micromolar NAD concentrations. Benzamides at nanomolar concentrations also activate mono-ADP-ribosylation of the enzyme, but at higher concentrations inhibit elongation at micromolar NAD as substrate. In nuclei, the enzyme molecule extensively auto-ADP-ribosylates itself, whereas histones are trans-ADP-ribosylated to a much lower extent. The unstable mono-ADP-ribose enzyme adduct represents an initiator intermediate in poly ADP-ribosylation.  相似文献   

18.
Liu D  Pozharski E  Lepore BW  Fu M  Silverman RB  Petsko GA  Ringe D 《Biochemistry》2007,46(37):10517-10527
As a mechanism-based inactivator of PLP-enzymes, (S)-4-amino-4,5-dihydro-2-thiophenecarboxylic acid (SADTA) was cocrystallized with Escherichia coli aspartate aminotransferase (l-AspAT) at a series of pH values ranging from 6 to 8. Five structural models with high resolution (1.4-1.85 A) were obtained for l-AspAT-SADTA complexes at pH 6.0, 6.5, 7.0, 7.5, and 8.0. Electron densities of the models showed that two different adducts had formed in the active sites. One adduct was formed from SADTA covalently linked to pyridoxal 5'-phosphate (PLP) while the other adduct was formed with the inhibitor covalently linked to Lysine246,1 the active site lysine. Moreover, there is a strong indication based on the electron densities that the occurrence of the two adducts is pH dependent. We conclude that SADTA inactivates l-AspAT via two different mechanisms based on the binding direction of the inactivator. Additionally, the structural models also show pH dependence of the protein structure itself, which provided detailed mechanistic implications for l-AspAT.  相似文献   

19.
The phosphorylated form of the N-terminal domain of enzyme I of the phosphoenolpyruvate:sugar phosphotransferase system of Escherichia coli has been investigated by one-bond and long-range 1H-15N correlation spectroscopy. The active site His 189 is phosphorylated at the Nepsilon2 position and has a pKa of 7.3, which is one pH unit higher than that of unphosphorylated His 189. Because the neutral form of unphosphorylated His 189 is in the Ndelta1-H tautomer, and its Nepsilon2 atom is solvent inaccessible and accepts a hydrogen bond from the hydroxyl group of Thr 168, both protonation and phosphorylation of His 189 must be accompanied by a change in the side-chain conformation of His 189, specifically from a chi(2) angle in the g+ conformer in the unphosphorylated state to the g- conformer in the phosphorylated state.  相似文献   

20.
The 2-ethoxycarbonyl-2-methyl-3,4-dihydro-2H-pyrrole-l-oxide (EMPO), an easily prepared pyrroline-N-oxide has been tested as a free radical scavenger. Spin adducts of superoxide, hydroxyl radical, and other free radicals were characterized in phosphate buffer at pH 7.0 and 5.6. At pH 7 in phosphate buffer, the EMPO/O(2)(-*) spin adduct was estimated to be about five times more persistent than its DMPO analogue. Furthermore, its decay does not produce the EMPO/HO&z.rad; adduct.  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