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.
The function of the binuclear Cu(A) center in cytochrome c oxidase (CcO) was studied using two Rhodobacter sphaeroides CcO mutants involving direct ligands of the Cu(A) center, H260N and M263L. The rapid electron-transfer kinetics of the mutants were studied by flash photolysis of a cytochrome c derivative labeled with ruthenium trisbipyridine at lysine-55. The rate constant for intracomplex electron transfer from heme c to Cu(A) was decreased from 40000 s(-1) for wild-type CcO to 16000 s(-1) and 11000 s(-1) for the M263L and H260N mutants, respectively. The rate constant for electron transfer from Cu(A) to heme a was decreased from 90000 s(-1) for wild-type CcO to 4000 s(-1) for the M263L mutant and only 45 s(-1) for the H260N mutant. The rate constant for the reverse reaction, heme a to Cu(A), was calculated to be 66000 s(-1) for M263L and 180 s(-1) for H260N, compared to 17000 s(-1) for wild-type CcO. It was estimated that the redox potential of Cu(A) was increased by 120 mV for the M263L mutant and 90 mV for the H260N mutant, relative to the potential of heme a. Neither mutation significantly affected the binding interaction with cytochrome c. These results indicate that His-260, but not Met-263, plays a significant role in electron transfer between Cu(A) and heme a.  相似文献   

2.
Saccharomyces cerevisiae phosphoenolpyruvate (PEP) carboxykinase catalyzes the reversible formation of oxaloacetate and adenosine triphosphate from PEP, adenosine diphosphate and carbon dioxide, and uses Mn(2+) as the activating metal ion. Comparison with the crystalline structure of homologous Escherichia coli PEP carboxykinase [Tari et al. Nature Struct. Biol. 4 (1997) 990-994] shows that Lys(213) is one of the ligands to Mn(2+) at the enzyme active site. Coordination of Mn(2+) to a lysyl residue is infrequent and suggests a low pK(a) value for the epsilon-NH(2) group of Lys(213). In this work, we evaluate the role of neighboring Phe(416) in contributing to provide a low polarity microenvironment suitable to keep the epsilon-NH(2) of Lys(213) in the unprotonated form. Mutation Phe416Tyr shows that the introduction of a hydroxyl group in the lateral chain of the residue produces a substantial loss in the enzyme affinity for Mn(2+), suggesting an increase of the pK(a) of Lys(213). A study of the effect of pH on K(m) for Mn(2+) indicate that the affinity of recombinant wild type enzyme for the metal ion is dependent on deprotonation of a group with pK(a) of 7.1+/-0.2, compatible with the low pK(a) expected for Lys(213). This pK(a) value increases at least 1.5 pH units upon Phe416Tyr mutation, in agreement with the expected effect of an increase in the polarity of Lys(213) microenvironment. Theoretical calculations of the pK(a) of Lys(213) indicate a value of 6.5+/-0.9, and it increases to 8.2+/-1.6 upon Phe416Tyr mutation. Additionally, mutation Phe416Tyr causes a loss of 1.3 kcal mol(-1) in the affinity of the enzyme for PEP, an effect perhaps related to the close proximity of Phe(416) to Arg(70), a residue previously shown to be important for PEP binding.  相似文献   

3.
The cytotoxic responses of midgut epithelial cells (MEC) from spruce budworm (SBW), gypsy moth (GM) and silkworm (SW) larvae were compared with the cytotoxic response of lepidopteran cell lines (SF-9, SE-1a, and CF-1) to CryIA toxins from Bacillus thuringiensis. The MEC from SBW, SW and GM had binding proteins for CryIA(a,b,c) toxins, whereas the lepidopteran cell lines had binding proteins for CryIA(c). Single MEC exposed to CryIA(a,b,c) toxins in a qualitative lawn assay were equally susceptible to the toxins with a threshold response at about 1ng. The cell lines were not susceptible to CryIA(a,b) toxins in the dose range tested, but had threshold responses for CryIA(c) of 3.4ng for SF-9, 50.2ng for SE-1a and 5.9ng for CF-1. In the quantitative Live/Dead assay, MEC were equally susceptible to CryIA(a,b,c) toxins with a threshold effect at about 1ng and a maximum effect at about 10ng. CF-1 was most sensitive to CryIA(c) with a threshold effect at 0.39ng and a maximal effect at about 1ng. In contrast, a 25-50 times greater dose of CryIA(a) or CryIA(b) was required to elicit a similar response as CryIA(c) for CF-1. SF-9 and SE-1a were most susceptible to CryIA(c) with a threshold effect observed at about 0.5ng and maximal effects at about 2ng. SF-9 cells have a threshold and maximum response to CryIA(a,b) of about 10ng and 20ng, respectively. SE-1a cells have a threshold and maximal response to CryIA(a,b) of 5ng and 10ng, respectively. Intact midgut epithelium exposed to CryIA(a,b,c) toxins had a threshold dose of 2ng for CryIA(b), 10-30ng for CryIA(a) and 2-30ng for CryIA(c). This study has shown that MEC are affected by a broader spectrum of toxins compared to the lepidopteran larvae and insect cell lines.  相似文献   

4.
5.
U B Goli  R E Galardy 《Biochemistry》1986,25(22):7136-7142
Five phosphorus-containing inhibitors of angiotensin converting enzyme were found to exhibit slow, tight-binding kinetics by using furanacryloyl-L-phenylalanylglycylglycine as substrate at pH 7.50 and T = 25 degrees C. Two of the inhibitors, (O-ethylphospho)-Ala-Pro (2) and (O-isopropylphospho)-Ala-Pro (3), are found to follow at minimum a two-step mechanism of binding (mechanism B) to the enzyme. This mechanism consists of an initial fast formation of a weaker enzyme-inhibitor complex (Ki = 130 nM for 2 and 180 nM for 3) followed by a slow reversible isomerization to a tighter complex with measurable forward (K3) and reverse (k4) rate constants (k3 = 4.5 X 10(-2) s-1 for 2 and 5.4 X 10(-2) s-1 for 3; k4 = 9.2 X 10(-3) s-1 for 2 and 3.5 X 10(-3) s-1 for 3). For the remaining three inhibitors, phospho-Ala-Pro (1), (O-benzyl-phospho)-Ala-Pro (4), and (P-phenethylphosphono)-Ala-Pro (5), a one-step binding mechanism (mechanism A) is observed under the conditions of the experiment. The second-order rate constants k1 (M-1 s-1) for the binding of these inhibitors to converting enzyme are found to have values more than 3 orders of magnitude lower than the diffusion-controlled limit for a bimolecular reaction involving the enzyme, viz., 3.9 X 10(5) for 1, 2.2 X 10(5) for 4, and 4.8 X 10(5) for 5.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
A mass spectrometer with a membrane-covered inlet was used to measure nitrogen fixation by following changes in the concentration of dissolved N(2) in a stirred suspension of the cyanobacterium Anabaena variabilis in an open system. The results showed a good fit to Michaelis-Menten kinetics with a K(m) for N(2) of 65 muM at 35 degrees C, corresponding to 0.121 atmosphere of N(2). Corresponding values for the K(m) for acetylene reduction were 385 muM (0.011 atmosphere at 35 degrees C). Comparison of the values of V(max) for N(2) uptake with those for the acetylene reduction assay under similar conditions gave an average value of 3.8 for the conversion factor between N(2) and C(2)H(2) reduction. Reduction of protons to hydrogen was completely inhibited at sufficiently high concentrations of C(2)H(2), but even at saturating N(2) concentrations, 1 mol of H(2) was produced for every mole of N(2) reduced. This explains the finding that the observed C(2)H(2)/N(2) ratio is higher than the value of 3 expected from the requirement for two electrons for acetylene reduction and six for nitrogen reduction. The results correlate well with a mechanism for N(2) reduction involving the equation: N(2) + 8H + 8e --> 2NH(3) + H(2) which gives a conversion factor between C(2)H(2) and N(2) of 4. It is proposed that, in general, 4 is a more appropriate value than 3 for the conversion factor.  相似文献   

7.
We synthesized methylvesamicol analogs 13-16 and investigated the binding characteristics of 2-[4-phenylpiperidino]cyclohexanol (vesamicol) and methylvesamicol analogs 13-16, with a methyl group introduced into the 4-phenylpiperidine moiety, to sigma receptors (sigma-1, sigma-2) and to vesicular acetylcholine transporters (VAChT) in membranes of the rat brain and liver. In competitive inhibition studies, (-)-o-methylvesamicol [(-)-OMV] (13) (Ki=6.7 nM), as well as (-)-vesamicol (Ki=4.4 nM), had a high affinity for VAChT. (+)-p-Methylvesamicol [(+)-PMV] (16) (Ki=3.0 nM), as well as SA4503 (Ki=4.4 nM), reported as a sigma-1 mapping agent for positron emission tomography (PET), had a high affinity for the sigma-1 receptor. The binding affinity of (+)-PMV (16) for the sigma-1 receptor (Ki=3.0 nM) was about 13 times higher than that for the sigma-2 (sigma-2) receptor (Ki=40.7 nM). (+)-PMV (16) (Ki=199 nM) had a much lower affinity for VAChT than SA4503 (Ki=50.2 nM) and haloperidol (Ki=41.4 nM). These results showed that the binding characteristics of (-)-OMV (13) to VAChT were similar to those of (-)-vesamicol and that (+)-PMV (16) bound to the sigma-1 receptor with high affinity. In conclusion, (-)-OMV (13) and (+)-PMV (16), which had a suitable structure, with a methyl group for labeling with 11C, may become not only a new VAChT ligand and a new type of sigma receptor ligand, respectively, but may also become a new target compound of VAChT and the sigma-1 receptor radioligand for PET, respectively.  相似文献   

8.
The metabolic physiology of the Crested Pigeon (Ocyphaps lophotes) and the Brush Bronzewing (Phaps elegans) is generally similar to that expected for birds of their size, but the Crested Pigeon has a number of characteristics which would aid survival in hot and dry regions. Body temperature increased similarly for the Crested Pigeon (from 38.8 degrees C to 41.5 degrees C) and the Brush Bronzewing (39.3 degrees C to 41.4 degrees C) over ambient temperatures (T(a)s) from 10 degrees C to 35 degrees C. Both species became hyperthermic (body temperature, T(b)>42 degrees C) at T(a)=45 degrees C. Basal metabolic rate of the Crested Pigeon (0.65 ml O(2) g(-1) h(-1) at 40 degrees C) was approximately 71% of that predicted for a columbid bird, while BMR of the Brush Bronzewing (0.87 ml O(2) g(-1) h(-1) at 20 degrees C to 40 degrees C) was approximately 102% of predicted. Total evaporative water loss increased exponentially with T(a) for both species, from <1 mg H(2)O g(-1) h(-1) at 10 degrees C to >12 mg H(2)O g(-1) h(-1) at 45 degrees C. It was similar and low for both species at T(a)<30 degrees C, but was higher for the Brush Bronzewing than the Crested Pigeon at T(a)>30 degrees C. Ventilatory minute volume matched oxygen consumption, such that oxygen extraction efficiency did not change with T(a) and was similar for both species (approximately 20%). Expired air temperature was considerably lower than T(b) for both species at T(a)<35 degrees C, potentially reducing respiratory water loss by approximately 65% at T(a)=10 degrees C to approximately 30% at T(a)=35 degrees C. Cutaneous evaporative cooling was significant for both species, with skin resistance decreasing as T(a) increased. The Crested Pigeon had a lower skin resistance than the Brush Bronzewing at T(a)=45 degrees C. The Brush Bronzewing had apparently reached its maximum cutaneous water loss at 30 degrees C and relied on panting to cool at higher T(a).  相似文献   

9.
Using an equilibrium dialysis technique, moles (Gamma(2)(1)) of cationic and anionic surfactants bound per kilogram of proteins of various types in aqueous media have been measured previously in this laboratory under different physicochemical conditions. From a thermodynamic analysis in the present paper, Gamma(2)(1) has been shown to be equal to the Gibbs relative excess of surfactant per kilogram of protein at a measured value of solute activity, a(2). The values of relative solvent excesses, Gamma(2)(1) (which are negative for surfactants) can be estimated from values of Gamma(2)(1) and a(2). Using the Gibbs-Duhem relationship for protein solution inside the dialysis bag and dialysate solutions respectively at equilibrium, an integrated expression for the standard free energy change, DeltaG(o) (in kilojoules per kilogram of protein for binding with ligand as a result of the change of a(2) from zero to unity) can be calculated from experimental data. The isopiestic vapour pressure technique was used extensively for evaluation of negative binding (-Gamma(2)(1)) of inorganic salts to proteins of different types for various values of a(2) of salts present in the bulk media. With some modifications of our derived equations for free energy of binding in such a system, DeltaG(o) has been evaluated for the change of mean activity of electrolyte from zero to unity in the rational scale. DeltaG(o) is positive since Gamma(2)(1) is negative and Gamma(2)(1) is positive for such ionic systems. DeltaG(o) in all cases, however, are expressed in terms of the standard state of reference of unit activity so that their magnitudes and sign can be related to the relative affinities of a solute for binding with proteins in aqueous media.  相似文献   

10.
CCA-adding enzyme builds the 3'-end CCA of tRNA without a nucleic acid template. The mechanism for the maintenance of fidelity during the CCA-adding reaction remains elusive. Here, we present almost a dozen complex structures of the class I CCA-adding enzyme and tRNA mini-helices (mini-D(73)N(74), mini-D(73)N(74)C(75) and mini-D(73)C(74)N(75); D(73) is a discriminator nucleotide and N is either A, G, or U). The mini-D(73)N(74) complexes adopt catalytically inactive open forms, and CTP shifts the enzymes to the active closed forms and allows N(74) to flip for CMP incorporation. In contrast, unlike the catalytically active closed form of the mini-D(73)C(74)C(75) complex, the mini-D(73)N(74)C(75) and mini-D(73)C(74)N(75) complexes adopt inactive open forms. Only the mini-D(73)C(74)U(75) accepts AMP to a similar extent as mini-D(73)C(74)C(75), and ATP shifts the enzyme to a closed, active form and allows U(75) to flip for AMP incorporation. These findings suggest that the 3'-region of RNA is proofread, after two nucleotide additions, in the closed, active form of the complex at the AMP incorporation stage. This proofreading is a prerequisite for the maintenance of fidelity for complete CCA synthesis.  相似文献   

11.
Clark DD  Boyd JM  Ensign SA 《Biochemistry》2004,43(21):6763-6771
2-[(R)-2-Hydroxypropylthio]ethanesulfonate (R-HPC) dehydrogenase (DH) catalyzes the reversible oxidation of R-HPC to 2-(2-ketopropylthio)ethanesulfonate (2-KPC) in a key reaction in the bacterial conversion of chiral epoxides to beta-keto acids. R-HPCDH is highly specific for the R-enantiomer of HPC, while a separate enzyme, S-HPCDH, catalyzes the oxidation of the corresponding S-enantiomer. In the present study, the features of substrate and enzyme imparting stereospecificity have been investigated for R-HPCDH. S-HPC was a substrate for R-HPCDH with a K(m) identical to that for R-HPC but with a k(cat) 600 times lower. Achiral 2-propanol and short-chain (R)- and (S)-2-alkanols were substrates for R-HPCDH. For (R)-alkanols, as the carbon chain length increased, K(m) decreased, with the K(m) for (R)-2-octanol being 1700 times lower than for 2-propanol. At the same time, k(cat) changed very little and was at least 90% lower than k(cat) for R-HPC and at least 22 times higher than k(cat) for S-HPC. (S)-2-Butanol and (S)-2-pentanol were substrates for R-HPCDH. The K(m) for (S)-2-butanol was identical to that for (R)-2-butanol, while the K(m) for (S)-2-pentanol was 7.5 times higher than for (R)-2-pentanol. Longer chain (S)-2-alkanols were sufficiently poor substrates for R-HPCDH that kinetic parameters could not be determined. Mutagenesis of C-terminal arginine residues of R-HPCDH revealed that R152 and R196 are essential for effective catalysis with the natural substrates R-HPC and 2-KPC but not for catalysis with 2-alkanols or ketones as substrates. Short-chain alkylsulfonates and coenzyme M (2-mercaptoethanesulfonate) were found to modify the kinetic parameters for 2-butanone reduction by R-HPCDH in a saturable fashion, with the general effect of increasing k(cat), decreasing K(m), and increasing the enantioselectivity of 2-butanone reduction to a theoretical value of 100% (S)-2-butanol. The modulating effects of ethanesulfonate and propanesulfonate provided thermodynamic binding constants close to K(m) for the natural substrates R-HPC and 2-KPC. The effects of alkylsulfonates on modulating the enantioselectivity and kinetic properties of R-HPCDH were abolished in R152A and R196A mutants but not in mutants of other C-terminal arginine residues. Collectively, the results suggest that interactions between the sulfonate of CoM and specific arginine residues are key to the enantioselectivity and catalytic efficiency of R-HPCDH. A model is proposed wherein sulfonate-arginine interactions within an alkylsulfonate binding pocket control the catalytic properties of R-HPCDH.  相似文献   

12.
We quantify Coulombic end effects (CEE) on oligocation-nucleic acid interactions at salt concentrations ([salt]) in the physiological range. Binding constants (K(obs); per site, at zero binding density) for the +8-charged C-amidated oligopeptide KWK6 and short single-stranded DNA oligonucleotides [dTpdT(|Z(D)|), where 6 < or = |Z(D)| < or = 22 is the number of DNA phosphates] were determined as a function of [salt] by fluorescence quenching. For the different DNA oligomers, K(obs) values are similar at high [salt], but diverge as [salt] decreases because -S(a)K(obs) identical with--partial partial differential ln K(obs)/ partial differential ln a+/- increases strongly with |Z(D)|. For binding of KWK6 near 0.1 M salt, -S(a)K(obs) is 5.5 +/- 0.2 for dT(pdT)22, 4.0 +/- 0.2 for dT(pdT)10 and 2.9 +/- 0.2 for dT(pdT)6, as compared with 6.5 +/- 0.3 for poly(dT). Similarly, at 0.1 M salt, K(obs) per site for poly(dT) exceeds K(obs) for dT(pdT)22 by 7-fold, for dT(pdT)10 by 50-fold and for dT(pdT)6 by 700-fold. We interpret the reductions in K(obs) and |S(a)K(obs)| with decreasing |Z(D)| as a significant CEE that causes binding to the terminal regions of a nucleic acid to be weaker and less salt dependent than interior binding. We analyze long oligonucleotide-KWK6 binding data in terms of a trapezoidal model for the local (axial) salt cation concentration on single-stranded DNA to estimate the size of the CEE to be at least seven phosphates on each end at 0.1 M salt.  相似文献   

13.
Intestinal absorption of strontium (Sr) in thirteen healthy adult German volunteers has been investigated by simultaneous oral and intravenous administration of two stable tracer isotopes, i.e. (84)Sr and (86)Sr. The measured Sr tracer concentration in plasma was analyzed using the convolution integral technique to obtain the intestinal absorption rate. The results showed that the Sr labeled in different foodstuffs was absorbed into the body fluids in a large range of difference. The maximum Sr absorption rates were observed within 60-120 min after administration. The rate of absorption is used to evaluate the intestinal absorption fraction, i.e. the f (1) value for various foodstuffs. The equivalent and effective dose coefficients for ingestion of (90)Sr were calculated using these f (1) values, and they were compared with those recommended by the International Commission on Radiological Protection (ICRP). The geometric and arithmetic means of the f (1) values are 0.38 and 0.45 associated with a geometric standard deviation and a standard deviation of 1.88 and 0.22, respectively. The 90% confidence interval of the f (1) values obtained in the present study ranges from 0.13 to 0.98. Expressed as the ratio of the 95 and 50% percentiles of the estimated probability, the uncertainty for the f (1) value corresponds to a factor of 2.58. The effective dose coefficients of (90)Sr after ingestion are 6.1 x 10(-9) Sv Bq(-1) for an f(1) value of 0.05, 1.0 x 10(-8) Sv Bq(-1) for 0.1, 1.9 x 10(-8) Sv Bq(-1) for 0.2, 2.8 x 10(-8) Sv Bq(-1) for 0.3, 3.6 x 10(-8) Sv Bq(-1) for 0.4, 5.3 x 10(-8) Sv Bq(-1) for 0.6, 7.1 x 10(-8) Sv Bq(-1) for 0.8, and 7.9 x 10(-8) Sv Bq(-1) for 0.9, respectively. Taking the effective dose coefficient of 2.8 x 10(-8) Sv Bq(-1) for an f (1) value of 0.3, which is recommended by the ICRP, as a reference, the effective dose coefficient of (90)Sr after ingestion varies by a factor of 2.8 when the f (1) value changes by a factor of 3, i.e. it decreases from 0.3 to 0.1 or increases from 0.3 to 0.9, respectively.  相似文献   

14.
Although portable instruments have been used in the assessment of sleep disturbance for patients with low back pain (LBP), the accuracy of the instruments in detecting sleep/wake episodes for this population is unknown. This study investigated the criterion validity of two portable instruments (Armband and Actiwatch) for assessing sleep disturbance in patients with LBP. 50 patients with LBP performed simultaneous overnight sleep recordings in a university sleep laboratory. All 50 participants were assessed by Polysomnography (PSG) and the Armband and a subgroup of 33 participants wore an Actiwatch. Criterion validity was determined by calculating epoch-by-epoch agreement, sensitivity, specificity and prevalence and bias- adjusted kappa (PABAK) for sleep versus wake between each instrument and PSG. The relationship between PSG and the two instruments was assessed using intraclass correlation coefficients (ICC 2, 1). The study participants showed symptoms of sub-threshold insomnia (mean ISI = 13.2, 95% CI = 6.36) and poor sleep quality (mean PSQI = 9.20, 95% CI = 4.27). Observed agreement with PSG was 85% and 88% for the Armband and Actiwatch. Sensitivity was 0.90 for both instruments and specificity was 0.54 and 0.67 and PABAK of 0.69 and 0.77 for the Armband and Actiwatch respectively. The ICC (95%CI) was 0.76 (0.61 to 0.86) and 0.80 (0.46 to 0.92) for total sleep time, 0.52 (0.29 to 0.70) and 0.55 (0.14 to 0.77) for sleep efficiency, 0.64 (0.45 to 0.78) and 0.52 (0.23 to 0.73) for wake after sleep onset and 0.13 (−0.15 to 0.39) and 0.33 (−0.05 to 0.63) for sleep onset latency, for the Armband and Actiwatch, respectively. The findings showed that both instruments have varied criterion validity across the sleep parameters from excellent validity for measures of total sleep time, good validity for measures of sleep efficiency and wake after onset to poor validity for sleep onset latency.  相似文献   

15.
For the determination of prostaglandin (PG) D(2) produced by cultured cells in response to external stimuli, immunological methods would be convenient and useful. However, PGD(2) is unstable under the physiological conditions, so that it has been difficult to get a specific antibody for the parent PGD(2). In an attempt to get a specific antibody for PGD(2), we tried to prepare monoclonal antibodies for 11-deoxy-11-methylene-PGD(2), a novel, chemically stable, isosteric analogue of PGD(2). We successfully cloned a hybridoma cell line secreting a monoclonal antibody reacting specifically with the parent PGD(2). To develop the enzyme-linked immunosorbent assay (ELISA) for PGD(2), the immobilized antigen using the stable PGD(2) derivative was immunoreacted in a competitive manner with the monoclonal antibody in presence of free PGD(2). The optimization of the assay provided a sensitive calibration curve for PGD(2) from 0.32 pg to 0.18 ng with a value of 7.6 pg at 50% displacement. PGD(2) was almost stable during the ELISA condition. The developed assay method was useful for applying to the direct determination of PGD(2) in the culture medium of mouse 3T3-L1 adipocytes. The incubation of PGD(2) in the maturation medium of adipocytes at 37 degrees C caused the chemical conversion into PGJ(2) derivatives. The conversion became more evident after 6 h of the incubation. These findings indicate the importance of considering the optimal time for collecting the samples to be determined for PGD(2) before the conversion starts to occur.  相似文献   

16.
Glutamate 47 is conserved in 1-aminocyclopropane-1-carboxylate (ACC) synthases and is positioned near the sulfonium pole of (S,S)-S-adenosyl-L-methionine (SAM) in the modeled pyridoxal phosphate quinonoid complex with SAM. E47Q and E47D constructs of ACC synthase were made to investigate a putative ionic interaction between Glu47 and SAM. The k(cat)/K(m) values for the conversion of (S,S)-SAM to ACC and methylthioadenosine (MTA) are depressed 630- and 25-fold for the E47Q and E47D enzymes, respectively. The decreases in the specificity constants are due to reductions in k(cat) for both mutant enzymes, and a 5-fold increase in K(m) for the E47Q enzyme. Importantly, much smaller effects were observed for the kinetic parameters of reactions with the alternate substrates L-vinylglycine (L-VG) (deamination to form alpha-ketobutyrate and ammonia) and L-alanine (transamination to form pyruvate), which have uncharged side chains. L-VG is both a substrate and a mechanism-based inactivator of the enzyme [Feng, L., and Kirsch, J. F. (2000) Biochemistry 39, 2436-2444], but the partition ratio, k(cat)/k(inact), is unaffected by the Glu47 mutations. ACC synthase primarily catalyzes the beta,gamma-elimination of MTA from the (R,S) diastereomer of SAM to produce L-VG [Satoh, S., and Yang, S. F. (1989) Arch.Biochem. Biophys. 271, 107-112], but catalyzes the formation of ACC to a lesser extent via alpha,gamma-elimination of MTA. The partition ratios for (alpha,gamma/beta,gamma)-elimination on (R,S)-SAM are 0.4, < or =0.014, and < or =0.08 for the wild-type, E47Q, and E47D enzymes, respectively. The results of these experiments strongly support a role for Glu47 as an anchor for the sulfonium pole of (S,S)-SAM, and consequently a role as an active site determinant of reaction specificity.  相似文献   

17.
A useful synthon to approach artificial phenylalanyl peptides in a [2 + 2 + 2] cycloaddition reaction, C(alpha,alpha)-dipropargylglycine (Dprg) is examined for its conformational preferences as a constrained residue. Crystal structure analysis and preliminary NMR results establish possible preference of the residue for folded (alpha) rather than extended (beta) region of the straight phi,psi conformational space. Boc-Dprg-L-Leu-OMe (1) displays two molecular conformations within the same crystallographic asymmetric unit, with Dprg in the alpha(R) or alpha(L) conformation, participating in a type I beta-turn or an alpha(L)-alpha(R)-type fold, in which Leu(2) assumes the alpha(R) conformation stereochemically favored for an L-chiral residue. Boc-Dprg-D-Val-L-Leu-OMe (2) displays a type I' beta-turn conformation in crystal, with both Dprg(1) and D-Val(2) assuming the alpha(L) conformation stereochemically favored for a D-chiral residue, with 4 --> 1 type hydrogen bond linking L-Leu(3) NH with Boc CO. NMR analysis using temperature variation, solvent titration, and a spin probe study suggests a fully solvent-exposed nature of Dprg NH, ruling out a fully extended C(5)-type conformation for this residue, and solvent sequestered nature of L-Leu(3) NH, suggesting possibility of a beta-turn due to Dprg assuming a folded conformation.  相似文献   

18.
Isolation and characterization of Z-DNA binding proteins from wheat germ   总被引:11,自引:0,他引:11  
E M Lafer  R Sousa  B Rosen  A Hsu  A Rich 《Biochemistry》1985,24(19):5070-5076
The preparation of a heterogeneous non-histone protein extract from wheat germ utilizing Br-poly(dG-dC).poly(dG-dC) (Z-DNA) affinity chromatography is described. The binding characteristics of antibodies against Z-DNA are used as a model system to define important criteria that the DNA binding behavior of a Z-DNA binding protein should display. We show that the wheat germ extract contains DNA binding proteins specific for left-handed Z-DNA by these criteria. The affinity of the proteins measured by competition experiments was approximately 10(5) greater for Br-poly(dG-dC).poly(dG-dC) (Z-DNA) than for poly(dG-dC).poly(dG-dC) (B-DNA). The affinity of the proteins for plasmid DNA increases with increasing negative superhelicity which is known to stabilize Z-DNA. The proteins are shown to compete with Z-DNA antibodies for binding to supercoiled plasmids. Finally, the affinity for two plasmids at a given superhelical density is greater for the plasmid containing an insert known to form Z-DNA than for a plasmid without the insert. The proteins exhibit a 2-3-fold greater affinity for stretches of (dC-dA)n.(dT-dG)n over stretches of (dG-dC)n.(dG-dC)n when both sequences are induced to form Z-DNA by supercoiling.  相似文献   

19.
20.
Lauer S  Goldstein B  Nolan RL  Nolan JP 《Biochemistry》2002,41(6):1742-1751
Cholera toxin entry into mammalian cells is mediated by binding of the pentameric B subunit (CTB) to ganglioside GM(1) in the cell membrane. We used flow cytometry to quantitatively measure in real time the interactions of fluorescently labeled pentameric cholera toxin B-subunit (FITC-CTB) with its ganglioside receptor on microsphere-supported phospholipid membranes. A model that describes the multiple steps of this mode of recognition was developed to guide our flow cytometric experiments and extract relevant equilibrium and kinetic rate constants. In contrast to previous studies, our approach takes into account receptor cross-linking, an important feature for multivalent interactions. From equilibrium measurements, we determined an equilibrium binding constant for a single subunit of FITC-CTB binding monovalently to GM(1) presented in bilayers of approximately 8 x 10(7) M(-1) while that for binding to soluble GM(1)-pentasaccharide was found to be approximately 4 x 10(6) M(-1). From kinetic measurements, we determined the rate constant for dissociation of a single site of FITC-CTB from microsphere-supported bilayers to be (3.21 +/- 0.03) x 10(-3) s(-1), and the rate of association of a site on FITC-CTB in solution to a GM(1) in the bilayer to be (2.8 +/- 0.4) x 10(4) M(-1) s(-1). These values yield a lower estimate for the equilibrium binding constant of approximately 1 x 10(7) M(-1). We determined the equilibrium surface cross-linking constant [(1.1 +/- 0.1) x 10(-12) cm(2)] and from this value and the value for the rate constant for dissociation derived a value of approximately 3.5 x 10(-15) cm(2) s(-1) for the forward rate constant for cross-linking. We also compared the interaction of the receptor binding B-subunit with that of the whole toxin (A- and B-subunits). Our results show that the whole toxin binds with approximately 100-fold higher avidity than the pentameric B-subunit alone which is most likely due to the additional interaction of the A(2)-subunit with the membrane surface. Interaction of cholera toxin B-subunit and whole cholera toxin with gangliosides other than GM(1) revealed specific binding only to GD1(b) and asialo-GM(1). These interactions, however, are marked by low avidity and require high receptor concentrations to be observed.  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