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.
In this work, we have determined the thermodynamic parameters of the reduction of four different variants of Thiobacillus versutus amicyanin by electrochemical techniques. In addition, the thermodynamic parameters were determined of the low-pH conformational change involving protonation of the C-terminal histidine ligand and the concomitant dissociation of this histidine from the Cu(I) ion. In these variants, the native C-terminal loop containing the Cys, His, and Met copper ligands has been replaced with the corresponding polypeptide segments of Pseudomonas aeruginosa azurin, Populus nigra plastocyanin, Alcaligenes faecalis S-6 pseudoazurin, and Thiobacillus ferrooxidans rusticyanin. For the reduction reaction, each loop invariably holds an entropic "memory" of the mother protein. The thermodynamics of the low-pH transition vary in a fashion that is species-dependent. When present, the memory effect again shows a large entropic component. In particular, loop elongation tends to favor the formation of the Cu(I)-His bond (hence disfavors His protonation, yielding lower pK(a) values) probably due to an increased flexibility of the loop in the reduced state. Overall, it appears that both reduction and low-pH transition are loop-responsive processes. The spacing between the ligands mostly affects the change in the conformational freedom that accompanies the reaction.  相似文献   

2.
J B Ames  M Ros  J Raap  J Lugtenburg  R A Mathies 《Biochemistry》1992,31(23):5328-5334
Time-resolved ultraviolet resonance Raman spectra of bacteriorhodopsin are used to study protein structural changes on the nanosecond and millisecond time scales. Excitation at 240 nm is used to selectively enhance vibrational scattering from tyrosine so that changes in its hydrogen bonding and protonation state can be examined. Both nanosecond and millisecond UV Raman difference spectra indicate that none of the tyrosine residues change ionization state during the BR----K and BR----M transitions. However, intensity changes are observed at 1172 and 1615 cm-1 in the BR----M UV Raman difference spectra. The 1615-cm-1 feature shifts down 25 cm-1 in tyrosine-d4-labeled BR, consistent with its assignment as a tyrosine vibration. The intensity changes in the BR----M UV Raman difference spectra most likely reflect an increase in resonance enhancement that occurs when one or more tyrosine residues interact more strongly with a hydrogen-bond acceptor in M412. The frequency of the v7a feature (1172 cm-1) in the BR----M UV Raman difference spectra supports this interpretation. The proximity of Tyr-185 and Asp-212 in the retinal binding pocket suggests that deprotonation of the Schiff base in M412 causes Tyr-185 to stabilize ionized Asp-212 by forming a stronger hydrogen bond.  相似文献   

3.
Peters RJ  Croteau RB 《Biochemistry》2002,41(6):1836-1842
Abietadiene synthase catalyzes two sequential, mechanistically distinct cyclization reactions in the formation of a mixture of abietadiene double bond isomers as the committed step in resin acid biosynthesis. Each reaction is carried out at a separate active site residing in a structurally distinct domain, and the reactions are kinetically separable. The first cyclization reaction is initiated by protonation of the terminal double bond of the universal diterpene precursor, geranylgeranyl diphosphate. The pH dependence of the overall reaction is consistent with an acid-base catalytic mechanism, and a divalent metal ion plays a role in this reaction probably by binding the diphosphate moiety to assist in positioning the substrate for catalysis. A putative active site for the protonation-initiated cyclization was defined by modeling abietadiene synthase and locating the DXDD motif previously shown to be involved in this reaction. A number of charged and aromatic residues, which are highly conserved in mechanistically related diterpene cyclases, line the putative active site. Alanine substitutions were made for each of these residues, as were asparagine and glutamate substitutions for the aspartates of the DXDD motif. Kinetic evaluation confirmed the involvement of most of the targeted residues in the reaction, and analysis of mutational effects on the pH-activity profile and affinity for a transition state analogue suggested specific roles for several of these residues in catalyzing the cyclization of geranylgeranyl diphosphate to (+)-copalyl diphosphate. A functional role was also suggested for the cryptic insertional element found in abietadiene synthase and other diterpene synthases that carry out similar protonation-initiated cyclizations.  相似文献   

4.
The structural alterations which occur in bacteriorhodopsin (bR) during dark adaptation (BR570----BR548) and the primary phototransition of the dark photocycle (BR548----KD610) have been investigated by Fourier transform infrared and UV difference spectroscopy. Possible contributions of tyrosine to the Fourier transform infrared difference spectra of these transitions were assigned by incorporating ring per-deuterated tyrosine into bR. Based on these data and UV difference measurements, we conclude that a stable tyrosinate exists in BR570 at physiological temperature and that it protonates during formation of BR548. A tyrosinate protonation has also been observed at low temperature during the primary phototransition of BR570 to the red-shifted photoproduct K630 (1). However, we now find that no tyrosine protonation change occurs during the primary phototransition of BR548 to the red-shifted intermediate KD610. Through analysis of bR containing isotopically labeled retinals, it was also determined that the chromophore of KD610 exits in a 13-trans, 15-cis configuration. On the basis of this evidence and previous studies on the structure of the chromophore in BR570, BR548, and K630, it appears that only the 13-trans,15-trans configuration of the protonated chromophore leads to a stable tyrosinate group. It is proposed that a tyrosinate residue is stabilized due to its interaction with the Schiff base positive charge in the BR570 chromophore. Isomerization of the chromophore about either the C13 = C14 or C = N bond disrupts this interaction causing a protonation of the tyrosinate.  相似文献   

5.
A stopped-flow technique was used to study the spectral changes occurring in bilirubin-albumin following a pH jump as well as following binding of bilirubin at 25 degrees C. The changes were studied in two wavelength ranges, 280-310 nm (tyrosine residues) and 400-510 nm (bound bilirubin). The changes were analyzed according to a scheme of consecutive unimolecular reactions. Spectral monitoring of a pH jump from 11.3 to 11.8 reveals that the bilirubin-albumin complex changes its structure in several steps. The UV absorption spectra show that 3.8 tyrosine residues ionize in the first step, 2.5 in the second, none in the third, and 0.8 in the fourth and following steps. The visible absorption spectrum of bound bilirubin changes in the second, third, and fourth steps. The bilirubin spectra of the different bilirubin-albumin complexes occurring in the transition show a common isosbestic point at 445 nm, indicating a change of the dihedral angle between the two bilirubin chromophores in a three-step reaction. It is suggested that 1 tyrosine residue is located close to the bilirubin site and is externalized in the second step. Bilirubin binding to albumin was monitored at two pH values, 11.3 and 11.8. At pH 11.3 the complex changes its structure in a three-step relaxation sequence. A change of the dihedral angle between the bilirubin chromophores can explain the spectral changes observed in the second and third relaxations. Protonation of 0.7 tyrosine residues occurs in the third relaxation, suggesting internalization of a tyrosine residue as a late consequence of bilirubin binding. At pH 11.8 a two-step relaxation sequence follows bilirubin binding. No tyrosine protonation occurs. Bilirubin is probably bound more superficially at pH 11.8 than at pH 11.3.  相似文献   

6.
Grzyska PK  Kim Y  Jackson MD  Hengge AC  Denu JM 《Biochemistry》2004,43(27):8807-8814
Dual-specificity phosphatases (DSPs) belong to the large family of protein tyrosine phosphatases that contain the active-site motif (H/V)CxxGxxR(S/T), but unlike the tyrosine-specific enzymes, DSPs are able to catalyze the efficient hydrolysis of both phosphotyrosine and phosphoserine/threonine found on signaling proteins, as well as a variety of small-molecule aryl and alkyl phosphates. It is unclear how DSPs accomplish similar reaction rates for phosphoesters, whose reactivity (i.e., pK(a) of the leaving group) can vary by more than 10(8). Here, we utilize the alkyl phosphate m-nitrobenzyl phosphate (mNBP), leaving-group pK(a) = 14.9, as a physiological substrate mimic to probe the mechanism and transition state of the DSP, Vaccinia H1-related (VHR). Detailed pH and kinetic isotope effects of the V/K value for mNBP indicates that VHR reacts with the phosphate dianion of mNBP and that the nonbridge phosphate oxygen atoms are unprotonated in the transition state. (18)O and solvent isotope effects indicate differences in the respective timing of the proton transfer to the leaving group and P-O fission; with the alkyl ester substrate, protonation is ahead of P-O fission, while with the aryl substrate, the two processes are more synchronous. Kinetic analysis of the general-acid mutant D92N with mNBP was consistent with the requirement of Asp-92 in protonating the ester oxygen, either in a step prior to significant P-O bond cleavage or in a concerted but asynchronous mechanism in which protonation is ahead of P-O bond fission. Collectively, the data indicate that VHR and likely all DSPs can match leaving-group potential with the timing of the proton transfer to the ester oxygen, such that diverse aryl and alkyl phosphoesters are turned over with similar catalytic efficiency.  相似文献   

7.
The effect of neurophysin-hormone interaction on the environment of the single tyrosine of bovine neurophysin (Tyr-49) and on that of the tyrosine of oxytocin and vasopressin was studied by fluorescence; tyrosine-free peptides were used to determine effects on Tyr-49, and acetylated neurophysin was used to determine effects on the hormone tyrosine. Binding increases the fluorescence intensity of Tyr-49 by 130% while the fluorescence of the hormone tyrosine is almost completely quenched. Correlation of these results with those obtained on binding oxytocin or vasopressin to native neurophysin indicates that in the hormone complexes less than half of the fluorescence of Tyr-49 is lost by F?rster energy transfer to the quenched hormone tyrosine. These results support spin-label studies in indicating that the distance between Tyr-49 and the tyrosine of hormone bound to the strong hormone binding site is greater than 5 A. In the absence of peptides, the fluorescence of Tyr-49 increases by 40% on lowering the pH from 6.2 to 2. Titration of the acid fluorescence transition in bovine neurophysins-I and -II, and in bovine neurophysin-II treated with carboxypeptidase B to remove the Arg-Arg-Val sequence at the carboxyl terminus, indicates that this transition is due to titration of a side-chain carboxyl with an intrinsic pK of 4.6. The effects of guanidine, glycerol, and disulfide cleavage on the magnitude of the acid transition indicate that the conformational information necessary for the transition resides within the amino acid sequence adjacent to Tyr-49. Accordingly, the fluorescence acid transition is attributed to decreased quenching by Glu-46 or Glu-47 upon protonation. Glycerol is shown to perturb the glutamate-tyrosine interaction in the absence of general conformational effects. Comparison of the fluorescence low-pH transition with that of the low-pH circular dichroism transition of nitrated neurophysins suggests that the fluorescence and CD transitions reflect related, but not necessarily identical, phenomena. In an appendix, evidence is presented which suggests that the products of carboxy-peptidase digestion of bovine neurophysin-II are the same as two minor bovine neurophysin components, one of which is neurophysin-C.  相似文献   

8.
The rate of iron release from the N-terminal and C-terminal monoferrictransferrins (FeN-transferrin and FeC-transferrin, respectively) has been studied at 37°C over the pH range 3.5–10.6 using EDTA as the accepting chelate. FeN-transferrin is the more facile except above pH 8.2. Plots of log10kobs against pH showed a deviation for both monoferrictransferrins between pH 5.6 and 6.0 and studies above and below this transition point indicated that iron release occures by different mechanisms. At low pH (< 5.6) the rate of release from FeN-transferrin is independent of the presence of EDTA or NaClO4, whereas Fec-transferrin shows a small but significant increase with increasing EDTA concentration. Rapid protonation of both monoferrictransferrins is followed by relatively slow release of Fe3+ which is subsequently chelated by EDTA. The slower release from Fec-transferrin is probably due to its greater binding strength for iron and the greater conformational stability of the C-terminal domain. Above pH 6.0 iron release from both monoferrictransferrins increases as the concentration of EDTA is increased. Direct attacks of EDTA probably occurs giving Fe-transferrin (HCO3). EDTA as a transition state or intermediate. The factors which may lead to the observed pH dependence of the rate include (i) protonation of groups directly bound to the iron, (ii) conformers which differ in degree of protonation and (iii) the degree of protonation of the attacking chelating agent. It is suggested that an increase in conformational fluctuations as the pH is lowered may play a very important role. Studies with differrictransferrin at pH 4.53 and 7.40 showed that when iron is released to EDTA the rate is independent of the occupancy of the other site; that is, the two sites are exhibiting non-co-operativity.  相似文献   

9.
The mechanism of acid and enzymatic hydrolysis of the N-glycosidic bond of AMP has been investigated by fitting experimentally observed kinetic isotope effects [Parkin, D. W., & Schramm, V. L. (1987) Biochemistry (preceding paper in this issue)] to calculated kinetic isotope effects for proposed transition-state structures. The sensitivity of the transition-state calculations was tested by "arying the transition-state structure and comparing changes in the calculated kinetic isotope effects with the experimental values of the isotope effect measurements. The kinetic isotope effects for the acid-catalyzed hydrolysis of AMP are best explained by a transition state with considerable oxycarbonium character in the ribose ring, significant bonding remaining to the departing adenine ring, participation of a water nucleophile, and protonation of the adenine ring. A transition-state structure without preassociation of the water nucleophile cannot be eliminated by the data. Enzymatic hydrolysis of the N-glycosidic bond of AMP by AMP nucleosidase from Azotobacter vinelandii was analyzed in the absence and presence of MgATP, the allosteric activator that increases Vmax approximately 200-fold. The transition states for enzyme-catalyzed hydrolysis that best explain the kinetic isotope effects involve early SN1 transition states with significant bond order in the glycosidic bond and protonation of the adenine base. The enzyme enforces participation of an enzyme-bound water molecule, which has weak bonding to C1' in the transition state. Activation of AMP nucleosidase by MgATP causes the bond order of the glycosidic bond in the transition state to increase significantly. Hyperconjugation in the ribosyl group is altered by enzymatic stabilization of the oxycarbonium ion. This change is consistent with the interaction of an amino acid on the enzyme. Together, these changes stabilize a carboxonium-like transition-state complex that occurs earlier in the reaction pathway than in the absence of allosteric activator. In addition to the allosteric changes that alter transition-state structure, the presence of other inductive effects that are unobserved by kinetic isotope measurements is also likely to increase the catalytic rate.  相似文献   

10.
The effects of removal of the tyrosine 96 hydrogen bond on the stability and conformational events of cytochrome P-450cam are presented in this communication. Hydrostatic pressure has been used as a tool to perturbe the structure leading to the formation of cytochrome P-420, an inactivated but soluble and undenatured form of the enzyme. We show that the spin transition of cytochrome P-450cam, which is known to be influenced by hydrostatic pressure, is affected by this single mutation. The free energy of stabilisation of native substrate-free cytochrome P-450cam is not affected by the removal of the tyrosine 96 hydrogen bond via mutagenesis to phenylalanine, whereas the substrate-bound protein shows a difference of 21 kJ/mol. These results, as well as an observed 110 ml/mol difference for the volume of the inactivation reaction between substrate-bound native and mutant proteins, have been interpreted in terms of a more hydrated heme pocket for the site-directed mutant at position 96 compared to the wild-type protein where camphor is tightly bound via the tyrosine 96 hydrogen bond and water excluded from the active site.  相似文献   

11.
McCain DF  Grzyska PK  Wu L  Hengge AC  Zhang ZY 《Biochemistry》2004,43(25):8256-8264
Protein tyrosine phosphatases (PTPs) constitute a large family of signaling enzymes that include both tyrosine specific and dual-specificity phosphatases that hydrolyze pSer/Thr in addition to pTyr. Previous mechanistic studies of PTPs have relied on the highly activated substrate p-nitrophenyl phosphate (pNPP), an aryl phosphate with a leaving group pK(a) of 7. In the study presented here, we employ m-nitrobenzyl phosphate (mNBP), an alkyl phosphate with a leaving group pK(a) of 14.9, which mimics the physiological substrates of the PTPs. We have carried out pH dependence and kinetic isotope effect measurements to characterize the mechanism of two important members of the PTP superfamily: Yersinia PTP (YopH) and Cdc25A. Both YopH and Cdc25A exhibit bell-shaped pH-rate profiles for the hydrolysis of mNBP, consistent with general acid catalysis. The slightly inverse (18)(V/K)(nonbridge) isotope effects (0.9999 for YopH and 0.9983 for Cdc25A) indicate a loose transition state with little nucleophilic participation for both enzymes. The smaller (18)(V/K)(bridge) primary isotope effects (0.9995 for YopH and 1.0012 for Cdc25A) relative to the corresponding isotope effects for pNPP hydrolysis suggest that protonation of the leaving group oxygen at the transition state by the general acid is ahead of P-O bond fission with the alkyl substrate, while general acid catalysis of pNPP by YopH is more synchronous with P-O bond fission. The isotope effect data also confirm findings from previous studies that Cdc25A utilizes general acid catalysis for substrates with a leaving group pK(a) of >8, but not for pNPP. Interestingly, the difference in the kinetic isotope effects for the reactions of aryl phosphate pNPP and alkyl phosphate mNBP by the PTPs parallels what is observed in the uncatalyzed reactions of their monoanions. In these reactions, the leaving group is protonated in the transition state, as is the case in PTP-catalyzed reactions. Also, the phosphoryl group in the transition states of the enzymatic reactions does not differ substantially from those of the uncatalyzed reactions. These results provide further evidence that these enzymes do not change the transition state but simply stabilize it.  相似文献   

12.
L C Kurz  C Frieden 《Biochemistry》1987,26(25):8450-8457
The 13C NMR spectra of [2-13C]- and [6-13C]purine ribosides have been obtained free in solution and bound to the active site of adenosine deaminase. The positions of the resonances of the bound ligand are shifted relative to those of the free ligand as follows: C-2, -3.7 ppm; C-6, -73.1 ppm. The binary complexes are in slow exchange with free purine riboside on the NMR time scale, and the dissociation rate constant is estimated to be 13.5 s-1 from the slow exchange broadening of the free signal. In aqueous solution, protonation of purine riboside at N-1 results in changes in 13C chemical shift relative to those of the free base as follows: C-2, -4.9 ppm; C-6, -7.9 ppm. The changes in chemical shift that occur when purine riboside binds to the enzyme indicate that the hybridization of C-6 changes from sp2 to sp3 in the binary complex with formation of a new bond to oxygen or sulfur. A change in C-2 hybridization can be eliminated as can protonation at N-1 as the sole cause of the chemical shift changes. The kinetic constants for the adenosine deaminase catalyzed hydrolysis of 6-chloro- and 6-fluoropurine riboside have been compared, and the reactivity order implies that carbon-halogen bond breaking does not occur in the rate-determining step. These observations support a mechanism for the enzyme in which formation of a tetrahedral intermediate is the most difficult chemical step. Enzymic stabilization of this intermediate may be an important catalytic strategy used by the enzyme to lower the standard free energy of the preceding transition state.  相似文献   

13.
Meng J  McKnight CJ 《Biochemistry》2008,47(16):4644-4650
Villin-type headpiece domains are compact F-actin-binding motifs that have been used extensively as a model system to investigate protein folding by both experimental and computational methods. Villin headpiece (HP67) harbors a highly helical, thermostable, and autonomously folding subdomain in the C terminus (HP35), and because of this feature, HP67 is usually considered to be composed of a N- and C-terminal subdomain. Unlike the C-terminal subdomain, the N-terminal subdomain consists mainly of loops and turns, and the folding is dependent upon the presence of the C-terminal subdomain. The pH sensitivity of this subdomain is thought to arise from, at least partially, protonation of H41 buried in the hydrophobic core. Substitution of this histidine with tyrosine, another permissive residue at this position for naturally occurring sequences, increases not only the pH stability of HP67 but also the thermal stability and the cooperativity of thermal unfolding over a wide pH range (0.9-7.5). The crystal structures of wild-type HP67 and the H41Y mutant, determined under the same conditions, indicate that the H41Y substitution causes only localized rearrangement around the mutated residue. The F-actin-binding motif remains essentially the same after the mutation, accounting for the negligible effect of the mutation on F-actin affinity. The hydrogen bond formed between the imidazole ring of H41 and the backbone carbonyl of E14 of HP67 is eliminated by the H41Y mutation, which renders the extreme N terminus of H41Y more mobile; the hydrogen bond formed between the imidazole ring of H41 and the backbone nitrogen of D34 is replaced with that between the hydroxyl group of Y41 and the backbone nitrogen of D34 after the H41Y substitution. The increased hydrophobicity of tyrosine compensates for the loss of hydrogen bonds in the extreme N terminus and accounts for the increased stability and cooperativity of the H41Y mutant.  相似文献   

14.
The oxidations of dichlorodihydrofluorescein and dihydrorhodamine by peroxynitrite are zero-order in the indicator between pH 3 and 10. The yield of the oxidized products, dichlorofluorescein and rhodamine, significantly increased at pH values>7, and the maximal molar yields were 0.47 +/- 0.04 mol rhodamine and 0.54 +/- 0.06 mol, dichlorofluorescein per mol peroxynitrite at pH 8.5. The increase in yield of oxidized products as a function of pH indicates that the peroxynitrite anion may form an adduct with the indicator, followed by protonation and oxidation of the indicator. Carbon dioxide decreased the yield of fluorescent products to about 5%, relative to peroxynitrite, and the rate of product formation is again zero-order in the indicator. Given this yield, it is proposed that nitrogen dioxide and trioxocarbonate (*1-) are the reactive species that oxidize the indicators.  相似文献   

15.
The hydroxyl group of Tyr alpha 42 in human hemoglobin forms a hydrogen bond with the carboxylate of Asp beta 99 which is considered to be one of the most important hydrogen bonds for stabilizing the "T-state." However, no spontaneous mutation at position 42 of the alpha subunit has been reported, and the role of the tyrosine has not been tested experimentally. Two artificial human mutant hemoglobins in which Tyr alpha 42 was replaced by phenylalanine or histidine were synthesized in Escherichia coli, and their proton NMR spectra were studied with particular attention to the hyperfine-shifted and hydrogen-bonded proton resonances. The site-directed mutagenesis of the Tyr alpha 42----Phe removes the hydrogen bond described above and prevents transition to the T-state so that the mutant Hb is rather similar to the "R-state" even when deoxygenated. On the other hand, the mutation from tyrosine to histidine causes less drastic structural changes, and its quaternary and tertiary structures are almost the same as native deoxy-Hb A. This may be attributed to the formation of a new hydrogen bond between His alpha 1(42) and Asp beta 2(99). These observations indicate that the hydrogen bond formed between Tyr alpha 42 and Asp beta 99 is required to convert unliganded Hb to the T-state.  相似文献   

16.
Photosystem II (PSII), the thylakoid membrane enzyme which uses sunlight to oxidize water to molecular oxygen, holds many organic and inorganic redox cofactors participating in the electron transfer reactions. Among them, two tyrosine residues, Tyr-Z and Tyr-D are found on the oxidizing side of PSII. Both tyrosines demonstrate similar spectroscopic features while their kinetic characteristics are quite different. Tyr-Z, which is bound to the D1 core protein, acts as an intermediate in electron transfer between the primary donor, P(680) and the CaMn? cluster. In contrast, Tyr-D, which is bound to the D2 core protein, does not participate in linear electron transfer in PSII and stays fully oxidized during PSII function. The phenolic oxygens on both tyrosines form well-defined hydrogen bonds to nearby histidine residues, His(Z) and His(D) respectively. These hydrogen bonds allow swift and almost activation less movement of the proton between respective tyrosine and histidine. This proton movement is critical and the phenolic proton from the tyrosine is thought to toggle between the tyrosine and the histidine in the hydrogen bond. It is found towards the tyrosine when this is reduced and towards the histidine when the tyrosine is oxidized. The proton movement occurs at both room temperature and ultra low temperature and is sensitive to the pH. Essentially it has been found that when the pH is below the pK(a) for respective histidine the function of the tyrosine is slowed down or, at ultra low temperature, halted. This has important consequences for the function also of the CaMn? complex and the protonation reactions as the critical Tyr-His hydrogen bond also steer a multitude of reactions at the CaMn? cluster. This review deals with the discovery and functional assignments of the two tyrosines. The pH dependent phenomena involved in oxidation and reduction of respective tyrosine is covered in detail. This article is part of a Special Issue entitled: Photosystem II.  相似文献   

17.
The X-ray crystal structure of recombinant wild-type azurin from Pseudomonas aeruginosa was determined by difference Fourier techniques using phases derived from the structure of the mutant His35Leu. Two data sets were collected from a single crystal of oxidized azurin soaked in mother liquor buffered at pH 5.5 and pH 9.0, respectively. Both data sets extend to 1.93 A resolution. The two pH forms were refined independently to crystallographic R-factors of 17.6% (pH 5.5) and 17.5% (pH 9.0). The conformational transition previously attributed to the protonation/deprotonation of residue His35 (pKa(red) = 7.3, pKa(ox) = 6.2), which lies in a crevice of the protein close to the copper binding site, involves a concomitant Pro36-Gly37 main-chain peptide bond flip. At the lower pH, the protonated imidazole N delta 1 of His35 forms a strong hydrogen bond with the carbonyl oxygen from Pro36, while at alkaline pH the deprotonated N delta 1 acts as an acceptor of a weak hydrogen bond from HN Gly37. The structure of the remainder of the azurin molecule, including the copper binding site, is not significantly affected by this transition.  相似文献   

18.
Molecular films from polymeric materials play an important role in basic research as well as in technology. We have synthesized the double chain diacetylenic ammonium lipid N-bis-(10, 12-pentacosadiinoyl)-oxy-ethyl)-N-methyl-amine (ONCO). We have characterized monomolecular films at the air/water interface by means of microfluorescence filmbalance techniques and by atomic force microscopy (AFM). ONCO forms stable monomolecular films that exhibit a fluid-solid phase transition with a transition enthalpy of 90 kJ/mol at 10 °C and neutral pH. The coexistence pressure was found to decrease with decreasing protonation, which is in contrast to the commonly found Coulomb mechanism. A change in the chain packing due to a different nitrogen bond angle is discussed as a possible mechanism. This model is cor-roborated by the finding that crystals at high and at low pH differ in their polymerization properties as measured by microfluorescence. Received: 9 October 1996 / Accepted: 13 March 1997  相似文献   

19.
The kinetics of the dissociation reaction under acidic conditions of the dimeric pig and chicken mitochondrial malate dehydrogenases (EC 1.1.1.37) have been studied. The dissociation of the pig enzyme is completely reversible. The pK for dissociation determined by light-scattering measurements agrees within experimental error with the pK value of 5.25 measured for a tyrosine-carboxylate pair. The rate constants for the dissociation reaction and for the protonation process of this tyrosine are in close agreement. Thus, the tyrosine-carboxylate pair can be used as indicator of the dissociation reaction. The dissociation of the chicken enzyme proceeds around pH 4.5 at a much lower rate. A true equilibrium between dimer and monomers is not found, since the monomer gradually unfolds at this pH. The monomers of both enzymes, pig and chicken mitochondrial malate dehydrogenase, show the same stability towards acid. The difference in stability of the dimeric forms, therefore, must be due to an altered subunit contact area.  相似文献   

20.
The influence of the local environment on the formation of a tyrosyl radical was investigated in modified photosynthetic reaction centers from Rhodobacter sphaeroides. The reaction centers contain a tyrosine residue placed approximately 10 A from a highly oxidizing bacteriochlorophyll dimer. Measurements by both optical and electron paramagnetic resonance spectroscopy revealed spectral features that are assigned as arising primarily from an oxidized bacteriochlorophyll dimer at low pH values and from a tyrosyl radical at high pH values, with a well-defined transition that occurred with a pK(a) of 6.9. A model based on the wild-type structure indicated that the Tyr at M164 is likely to form a hydrogen bond with His M193 and to interact weakly with Glu M173. Substitution of Tyr or Glu for His at M193 increased the pK(a) for the transition from 6.9 to 8.9, while substitution of Gln for His M193 resulted in a higher pK(a) value. Substitution of Glu M173 with Gln resulted in loss of the partial formation of the tyrosyl that occurs in the other mutants at low pH values. The results are interpreted in terms of the ability of the residues to act as proton acceptors for the oxidized tyrosine, with the pK(a) values reflecting those of either the putative proton acceptor or the tyrosine, in accord with general models of amino acid radicals.  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