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.
This work is devoted to the problem connected with rhodopsin activation. Electrostatic forces involved in photoisomerization of retinal are considered. It is suggested that the repulsion force and rotating moment between electric dipole moments of methyl groups on the C5 and C13 atoms of retinal can promote isomerization upon light absorption because the pi-pi* transition is accompanied by a decrease in the potential barrier for torsional rotations around the C11-C12 bond.  相似文献   

2.
Modified trypsin kallikrein inhibitor (I*), with the reactive-site peptide bond Lys-15--Ala-16 split, reacts with alpha-chymotrypsin (E) via an intermediate X to the stable tetrahedral complex C:E + I in equilibrium X leads to C. Formation X constitutes a fast pre-equilibrium (equilibrium constant Kx = 7 X 10(-5) M, association rate constant kx = 4 X 10(3)M-1s-1) to the slow reaction X leads to C (rate constant kc = 2 X 10(-3) s-1), all values at pH 7.5. No intermediate X is observed when alpha-chymotrypsin reacts with I*-OMe in which the carboxyl group of Lys-15 is esterified by methanol. This observation as well as the different pH dependence of the overall association rate constants in the case of I* and I*-OMe indicate tha formation of X precedes formation of the acyl enzyme in the catalytic pathway. The data are compared to the similar results obtained with beta-trypsin and I* or I*-OMe.  相似文献   

3.
Dong J  Lu X  Wei Y  Luo L  Dunaway-Mariano D  Carey PR 《Biochemistry》2003,42(31):9482-9490
4-Chlorobenzoyl-coenzyme A (4-CBA-CoA) dehalogenase catalyzes the hydrolytic dehalogenation of 4-CBA-CoA to 4-hydroxybenzoyl-CoA by using an active site aspartate as the nucleophile. Formation of the corresponding Meisenheimer complex (EMc) is followed by chloride ion expulsion which forms the arylated intermediate (EAr). This is then hydrolyzed to the product. In this paper, we explore the relationship between active site polarizing forces acting on the benzoyl carbonyl and the rate of formation of the Meisenheimer complex. The polarizing forces at the C[double bond]O group were modulated by introducing site-selected mutations (A112V, Y65D, G113A, G113S, G113N, and F64P), near the C[double bond]O binding site. Using either the substrate, 4-CBA-CoA, or the substrate analogue, 4-methylbenzoyl-CoA (4-MBA-CoA), Raman difference spectroscopy provided the position of the C[double bond]O stretching frequency (nu(C)[double bond](O)) for a total of 10 enzyme-ligand complexes. In turn, the values of the C[double bond]O frequencies could be converted to differences in effective hydrogen bonding strengths between members of the series, based on earlier model studies [Clarkson, J., Tonge, P. J., Taylor, K. L., Dunaway-Mariano, D., and Carey, P. (1997) Biochemistry 36, 10192-10199]. Catalysis in the F64P, G113A, G113S, and G113N dehalogenase mutants was very slow with k(cat) values ranging from 8 x 10(-3) to 7.6 x 10(-6) s(-1). The EAr intermediate did not accumulate to a detectable level on these enzymes during a single turnover. Catalysis in the Y65D and A112V dehalogenase mutants were almost as efficient as catalysis in wild-type dehalogenase with k(cat) values of 0.1-0.6 s(-1). In wild-type dehalogenase, 22% of the bound substrate accumulated as the EAr intermediate during a single turnover (k(obs) for EAr formation = 24 s(-(1)); in the Y65D mutant, the level of accumulation is 17% (k(obs) for EAr formation = 3 s(-1)), and in the A112V mutant, the level is 23% (k(obs) for EAr formation = 17 s(-1)). The k(obs) for EAr formation in wild-type dehalogenase and the more active dehalogenase mutants (Y65D and A112V) was taken to be an estimate of the k for EMc formation, and the k(obs) for EP formation in a single turnover was taken to be an estimate of the k for EMc formation in the severely impaired mutants (F64P, G113A, G113S, and G113N). A plot of the log k(obs) for EMc formation versus the C[double bond]O stretching frequency of bound 4-CBA-CoA (or 4-MBA-CoA) is a straight line (R(2) = 0.9584). Throughout the series, nu(C)[double bond](O) varied by 61 cm(-1), corresponding to the change in hydrogen bonding enthalpy of 67 kJ/mol. The results show that changes in polarizing forces at the benzoyl carbonyl are transmitted to the benzoyl (4) position and correlate with the rate of aromatic nucleophilic addition five chemical bonds away. Interestingly, the relationship between effective polarizing forces and reactivity seen here for dehalogenase is similar to that reported for the addition-elimination reaction involving the hydrolysis of a series of acyl serine proteases.  相似文献   

4.
Planar 2(5H)-furanones substituted at C4 with a chiral pyrrolidinyl group show CD spectra which are apparently due to the distortion of the C4-N1 bond of sp2 character from the plane defined by the 2(5H)-furanone ring atoms and/or due to the presence of substituents in the pyrrolidine ring. This is a new, previously not encountered structural factor determining the chiroptical properties of 2(5H)-furanones and emerging from the analysis of X-ray diffraction data and quantum mechanical DFT computations. In the presence of a C5 pseudoaxial substituent in the furanone ring, the sign of the furanone n-pi* and pi-pi* transition Cotton effects is determined primarily by the previously postulated allylic helicity rule.  相似文献   

5.
Drohat AC  Stivers JT 《Biochemistry》2000,39(39):11865-11875
Uracil DNA glycosylase (UDG) cleaves the glycosidic bond of deoxyuridine in DNA using a hydrolytic mechanism, with an overall catalytic rate enhancement of 10(12)-fold over the solution reaction. The nature of the enzyme-substrate interactions that lead to this large rate enhancement are key to understanding enzymatic DNA repair. Using (1)H and heteronuclear NMR spectroscopy, we have characterized one such interaction in the ternary product complex of Escherichia coli UDG, the short (2.7 A) H bond between His187 N(epsilon)(2) and uracil O2. The H bond proton is highly deshielded at 15.6 ppm, indicating a short N-O distance and exhibits a solvent exchange rate that is 400- and 10(5)-fold slower than free imidazole at pH 7.5 and pH 10, respectively. Heteronuclear NMR experiments at neutral pH show that this H bond involves the neutral imidazole form of His187 and the N1-O2 imidate form of uracil. The excellent correspondence of the pK(a) for the disappearance of the H bond (pK(a) = 6.3 +/- 0.1) with the previously determined pK(a) = 6.4 for the N1 proton of enzyme-bound uracil indicates that the H bond requires negative charge on uracil O2 [Drohat, A. C., and Stivers, J. T. (2000) J. Am. Chem. Soc. 122, 1840-1841]. Although the above characteristics suggest a short strong H bond, the D/H fractionation factor of phi = 1.0 is more typical of a normal H bond. This unexpected observation may reflect a large donor-acceptor pK(a) mismatch or the net result of two opposing effects on vibrational frequencies: decreased N-H bond stretching frequencies (phi < 1) and increased bending frequencies (phi > 1) relative to the O-H bonds of water. The role of this H bond in catalysis by UDG and several approaches to quantify the H bond energy are discussed.  相似文献   

6.
The Cu(II) sites of native, azido- and cyano-derivatives of bovine superoxide dismutase (superoxide:superoxide oxidoreductase, EC 1.15.1.1) have been examined by electron-nuclear double resonance (ENDOR). The ENDOR spectrum of the native protein taken at the g parallel extreme shows resolved structure due to the directly coordinated N-atoms of the histidine ligands. These spectra are too complex for interpretation but suggest inequivalent coupling between the electronic spin and the four ligand N-atoms. By contrast, the azido protein reveals one type of nitrogen with well-resolved hyperfine and quadrupole splittings (Azz = 37.9 +/- 1 MHz, Pzz = 1.54 +/- 0.02 MHz), and the cyano from reveals one well-resolved set of nitrogen lines (Azz = 47.8 +/- 0.4 MHz, Pzz = 1.62 +/- 0.01 MHz) and one type of partially resolved nitrogen (Azz = 37.0 +/- 1 MHz). The cyano form also reveals a complex spectrum in the low-frequency domain (1-10 MHz). Through isotopic substitution and computer stimulation, the spectrum is shown to be a composite of the ENDOR from the remote imidazole nitrogens and the cyanide nitrogen. The component of the hyperfine constant perpendicular to the C14N bonds axis is A perpendicular N = 3.9 +/- 0.3 MHz and along the bond axis is A perpendicular N approximately equal to 5.7 MHz. The quadrupole interaction appears to be greatest along the CN axis with Qz'z' = 1.0 +/- 0.1 MHz and Qx'x'y'y' approximately 0. Based on an analysis of the hyperfine and quadrupole interactions seen at two extremes of the electron paramagnetic spectrum, we propose a square-planar arrangement of three imidazole nitrogen and one CN- carbon around the copper. Within this plane two imidazole nitrogens are strongly coupled and magnetically equivalent, the third is inequivalent (slightly weaker hyperfine interactions) and forms a trans relationship with the cyanide. This model is consistent with other observations on the cyano-derivative.  相似文献   

7.
The Met16Phe mutant of the type 1 copper protein pseudoazurin (PACu), in which a phenyl ring is introduced close to the imidazole moiety of the His81 ligand, has been characterized. NMR studies indicate that the introduced phenyl ring is parallel to the imidazole group of His81. The mutation has a subtle effect on the position of the two S(Cys)-->Cu(II) ligand-to-metal charge transfer bands in the visible spectrum of PACu(II) and a more significant influence on their intensities resulting in a A(459)/A(598) ratio of 0.31 for Met16Phe as compared to a A(453)/A(594) ratio of 0.43 for wild-type PACu(II) at pH 8. The electron paramagnetic resonance spectrum of the Met16Phe variant is more axial than that of the wild-type protein, and the resonance Raman spectrum of the mutant exhibits subtle differences. A C(gamma)H proton of Met86 exhibits a much smaller hyperfine shift in the paramagnetic (1)H NMR spectrum of Met16Phe PACu(II) as compared to its position in the wild-type protein, which indicates a weaker axial Cu-S(Met86) interaction in the mutant. The Met16Phe mutation results in an approximately 60 mV increase in the reduction potential of PACu. The pK(a) value of the ligand His81 decreases from 4.9 in wild-type PACu(I) to 4.5 in Met16Phe PACu(I) indicating that the pi-pi contact with Phe16 stabilizes the Cu-N(His81) interaction. The Met16Phe variant of PACu has a self-exchange rate constant at pH 7.6 (25 degrees C) of 9.8 x 10(3) M(-)(1) s(-)(1) as compared to the considerably smaller value of 3.7 x 10(3) M(-)(1) s(-)(1) for the wild-type protein under identical conditions. The enhanced electron transfer reactivity of Met16Phe PACu is a consequence of a lower reorganization energy due to additional active site rigidity caused by the pi-pi interaction between His81 and the introduced phenyl ring.  相似文献   

8.
Complete chemical selectivity (i.e., chemospecificity) has been achieved in the homogeneous deuteration of C5-C6 and endocyclic C10-C11 prostaglandin double bonds without arrangement or partial reduction of C13-C14- or C8-C12 double bonds. The homogeneous deuteration reaction utilizes protection of the C13-C14 double bond as the C15O-silyl ether and protection of the carboxyl group as the methyl ester prior to reduction under molecular deuterium with tris(triphenylphosphine)chlorohodium (I) (Wilkinson's catalyst) in 60:40 acetone:benzene at 25 degrees C. The reaction has been used to prepare six specifically deuterated prostaglandin: 5,6-dideuterio-PGE1 alpha 5,6-dideuterio-PGE1, 5,6-dideuterio-PGB1, 3,3,4,4,5,6-hexadeuterio-PGF1 alpha, 5,6,10,11-tetradeuterio-11-deoxy-PGE1, and 10, 11-dideuterio-11-deoxy-PGE1.  相似文献   

9.
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   

10.
E P Lennette  B V Plapp 《Biochemistry》1979,18(18):3933-3938
The reaction of the imidazole group of histidine hydantoin with bromoacetate was studied as a model for carboxymethylation of histidine residues in proteins. pK values of 6.4 and 9.1 (25 degrees C) and apparent heats of ionization of 7.8 and 8.7 kcal/mol were determined for the imidazole and hydantoin rings, respectively. At pH values corresponding to the isoelectric points for histidine hydantoin, the rates of carboxymethylation at 12, 25, 37, and 50 degrees C were determined; the modified hydantoins were hydrolyzed to the corresponding histidine derivatives for quantitative amino acid analysis. At pH 7.72 and 25 degrees C, the imidazole tele-N was alkylated (k = 3.9 X 10(-5) M-1 s-1) twice as fast as the pros-N. The monocarboxymethyl derivatives were carboxymethylated at the same rate at the pros-N (k = 2.1 X 10(-5) M-1 s-1) but 3 times faster at the tele-N (k = 11 X 10(-5) M-1 s-1). The enthalpies of activation determined for carboxymethylation of the imidazole ring and its monocarboxymethyl derivatives were similar (15.9 +/- 0.7 kcal/mol). delta S for the four carboxymethylations was -25 +/- 2 eu. The electrostatic component of delta S (delta S es) was calculated from the influence of the dielectric constant on the reaction rate at 25 degrees C. delta S es was slightly negative (-4 +/- 1 eu) for mono- or dicarboxymethylations, indicating some charge separation in the transition state. The nonelectrostatic entropy of activation was -21 +/- 2 eu for all four carboxymethylations.  相似文献   

11.
The photochemistry of three structurally very similar triphenylmethylsilanes 1, 2, 3 [p-X-C(6)H(4)-CPh(2)-SiMe(3): X = PhCO, 1; H, ; Ph(OCH(2)CH(2)O)C, 3] is described by means of 248 and 308 nm nanosecond laser flash photolysis (ns-LFP), femtosecond LFP, EPR spectroscopy, emission spectroscopy (fluorescence, phosphorescence), ns-pulse radiolysis (ns-PR), photoproduct analysis studies in MeCN, and X-ray crystallographic analysis of the two key-compounds 1 and 2. The photochemical behavior of 1, 2 and 3 is discussed and compared with that of a fourth one, 4, bearing on the p-position an amino group (X = Me(2)N) and whose detailed photochemistry we reported earlier (J. Org. Chem., 2000, 65, 4274-4280). Silane 1 undergoes on irradiation with 248 and 308 nm laser light a fast photodissociation of the C-Si bond giving the p-(benzoyl)triphenylmethyl radical (1*) with a rate constant of k(diss)= 3 x 10(7) s(-1). The formation of 1* is a one-quantum process and takes place via the carbonyl triplet excited state with high quantum yield (Phi(rad)= 0.9); the intervention of the triplet state is clearly demonstrated through the phosphorescence spectrum and quenching experiments with ferrocene (k(q)= 9.3 x 10(9) M(-1) s(-1)), Et(3)N (1.1 x 10(9) M(-1) s(-1)), and styrene (3.1 x 10(9) M(-1) s(-1)) giving quenching rate constants very similar to those of benzophenone. For comparative reasons radical 1* was generated independently from p-(benzoyl)triphenylmethyl bromide via pulse radiolysis in THF and its absorption coefficient at lambda(max)= 340 nm was determined ([epsilon]= 27770 M(-1) cm(-1)). We found thus that the p-PhCO-derivative 1 behaves similar to the p-Me(2)N one (the latter giving the p-(dimethylamino)triphenylmethyl radical with Phi(rad)= 0.9), irrespective of their completely different ground state electronic properties. In contrast, compounds 2, 3 that bear only the aromatic chromophore give by laser or lamp irradiation both, (i) radical products [Ph(3)C* and p-Ph(OCH(2)CH(2)O)C-C(6)H(4)-C(*)Ph(2), respectively] after dissociation of the central C-Si bond (Phi(rad)= 0.16), and (ii) persistent photo-Fries rearrangement products (of the type of 5-methylidene-6-trimethylsilyl-1,3-cyclohexadiene) absorbing at 300-450 nm and arising from a 1,3-shift of the SiMe(3) group from the benzylic to the ortho-position of the aromatic ring (Phi approximately 0.85 for 2). Using fs-LFP on 2 we showed that the S(1) state recorded at 100 fs after the pulse decays on a time scale of 500 fs giving Ph(3)C* through C-Si bond dissociation. In a second step and within the next 10 ps trityl radicals either escape from the solvent cage (the quantum yield of Ph(3)C* formation Phi(rad)= 0.16 was measured with ns-LFP), or undergo in-cage recombination to photo-Fries products. Thus, singlet excited states (S(1)) of the aromatic organosilanes (2, 3) prefer photo-Fries rearrangement products, while triplet excited states (1, 4) favor free radicals. Both reactions proceed via a common primary photodissociation step (C-Si bond homolysis) and differentiate obviously in the multiplicity of the resulting geminate radical pairs; singlet radical pairs give preferably photo-Fries products following an in-cage recombination, while triplet radical pairs escape the solvent cage (MeCN). The results demonstrate the crucial role which is played by the chromophore which prescribes in a sense, (i) the multiplicity of the intervening excited state and consequently that of the resulting geminate radical pair, and (ii) the dominant reaction path to be followed: the benzophenone- and anilino-chromophore present in silanes 1 and 4, respectively, impose effective intersystem crossing transitions (k(isc)= 10(11) s(-1) and 6 x 10(8) s(-1), respectively) leading to triplet states and finally to free radical products, while the phenyl chromophore in 2 and 3, possessing ineffective isc (k(isc)= 6 x 10(6) s(-1)) leads to photo-Fries product formation via the energetic high lying S(1) state [approximately 443 kJ mol(-1)(106 kcal mol(-1))].  相似文献   

12.
The chiroptical properties of the cyclic dipeptides cyclo-L -alanyl-L -histidine and cyclo-L -histidinyl-L -histidine have been investigated as a function of molecular conformation. The rotatory strengths of the n-π* transitions of the peptide chromophores and the lowest energy π-π* transitions of the imidazole chromophores have been calculated as a function of the angle of fold of the cyclic dipeptide group and the dihedral angles χ1 and χ2 of the amino acid side chains. The results of this investigation are consistent with the preferred position of the dihedral angle χ1 occurring near 60° in the free base form of cyclo-L -alanyl-L -histidine, and near 180° when the imidazole side chain is protonated. Furthermore, in the case of the free base form of the imidazole group, it is possible that the tautomeric isomer in which Nε is protonated may be more prevalent than the isomer in which Nδ is protonated.  相似文献   

13.
14.
A 10-mer random peptide library displayed on filamentous bacteriophage was used to determine the molecular basis of the interaction between the monoclonal anti-colicin A antibody 1C11 and its cognate epitope. Previous studies established that the putative epitope recognized by 1C11 antibody is composed of amino acid residues 19-25 (RGSGPEP) of colicin A. Using the phage display technique it was confirmed that the epitope of 1C11 antibody was indeed restricted to residues 19-25 and the consensus motif RXXXPEP was identified. Shorter consensus sequences (RXXPEP, RXXEP, KXXEP) were also selected. It was also demonstrated that the disulfide bond found in one group of the selected peptides was crucial for 1C11 antibody recognition. It was shown that cyclization of the peptides by disulfide bond formation could result in a structure that mimics the natural epitope of colicin A.  相似文献   

15.
Circular dichroism studies of seven helical oligopeptides containing alpha-aminoisobutyric acid (Aib) in methanol and trifluoroethanol (TFE) solutions are reported. Peptides ranging from 10 to 21 residues in length have been examined. In all cases distinct negative CD bands characteristic of helical peptides are obtained at approximately 220 and 205 nm corresponding to the n-pi and pi-pi transitions, respectively. The ratio R = [theta] pi-pi is less than 1.0 for all peptides studied. Using crystal structure and n.m.r. results for a 10 residue 3(10) helical peptide and literature values for an alpha-helical 11-residue peptide, it is shown that both helical conformations yield R values of approximately 0.8 in alcoholic solvents. The CD data are considered in the light of 1H n.m.r. studies on these oligopeptides. The results suggest that 3(10) and alpha-helical conformations cannot be distinguished by CD methods.  相似文献   

16.
Epoxyeicosatrienoic acids (EETs) are endothelium-derived eicosanoids that activate potassium channels, hyperpolarize the membrane, and cause relaxation. We tested 19 analogs of 14,15-EET on vascular tone to determine the structural features required for activity. 14,15-EET relaxed bovine coronary arterial rings in a concentration-related manner (ED(50) = 10(-6) M). Changing the carboxyl to an alcohol eliminated dilator activity, whereas 14,15-EET-methyl ester and 14,15-EET-methylsulfonimide retained full activity. Shortening the distance between the carboxyl and epoxy groups reduced the agonist potency and activity. Removal of all three double bonds decreased potency. An analog with a Delta8 double bond had full activity and potency. However, the analogs with only a Delta5 or Delta11 double bond had reduced potency. Conversion of the epoxy oxygen to a sulfur or nitrogen resulted in loss of activity. 14(S),15(R)-EET was more potent than 14(R),15(S)-EET, and 14,15-(cis)-EET was more potent than 14,15-(trans)-EET. These studies indicate that the structural features of 14,15-EET required for relaxation of the bovine coronary artery include a carbon-1 acidic group, a Delta8 double bond, and a 14(S),15(R)-(cis)-epoxy group.  相似文献   

17.
This study presents the first detailed examination by resonance Raman (RR) spectroscopy of the rates of solvent exchange for the C5 and C3 positions of the TPQ cofactor in several wild-type copper-containing amine oxidases and mutants of the amine oxidase from Hansenula polymorpha (HPAO). On the basis of crystal structure analysis and differing rates of C5 [double bond] O and C3 [bond] H exchange within the enzyme systems, but equally rapid rates of C5 [double bond] O and C3 [bond] H exchange in a TPQ model compound, it is proposed that these data can be used to determine the TPQ cofactor orientation within the active site of the resting enzyme. A rapid rate of C5 [double bond] O exchange (t(1/2) < 30 min) and a slow (t(1/2) = 6 h) to nonexistent rate of C3 [bond] H exchange was observed for wild-type HPAO, the amine oxidase from Arthrobacter globiformis, pea seedling amine oxidase at pH 7.1, and the E406Q mutant of HPAO. This pattern is ascribed to a productive TPQ orientation, with the C5 [double bond] O near the substrate-binding site and the C3 [bond] H near the Cu. In contrast, a slow rate of C5 [double bond] O exchange (t(1/2) = 1.6-3.3 h) coupled with a fast rate of C3 [bond] H exchange (t(1/2) < 30 min) was observed for the D319E and D319N catalytic base mutants of HPAO and for PSAO at pH 4.6 (t(1/2) = 4.5 h for C5 [double bond] O exchange). This pattern identifies a flipped orientation, involving 180 degrees rotation about the C alpha-C beta bond, which locates the C3 [bond] H near the substrate-binding site and the C5 double bond] O near the Cu. Finally, fast rates of both C5 [double bond] O and C3 [bond] H exchange (t(1/2) < 30 min) were observed for the amine oxidase from Escherichia coli and the N404A mutant of HPAO, suggesting a mobile cofactor, with multiple TPQ orientations between productive and flipped. These results demonstrate that opposing sides of the TPQ ring possess different degrees of solvent accessibility and that the rates of C5 [double bond] O and C3 [bond] H exchange can be used to predict the TPQ cofactor orientation in the resting forms of these enzymes.  相似文献   

18.
M N Boyden  S A Asher 《Biochemistry》2001,40(45):13723-13727
We used UV resonance Raman spectroscopy (UVRR) excited within the peptide bond pi --> pi* electronic transitions and within the aromatic amino acid pi --> pi* electronic transitions to examine the temperature dependence of the solution conformation of betanova, a 20-residue beta-sheet polypeptide [Kortemme, T., Ramirez-Alvarado, M., and Serrano, L. (1998) Science 281, 253-256]. The 206.5 nm excited UVRR enhances the amide vibrations and demonstrates that betanova has a predominantly beta-sheet structure between 5 and 82 degrees C. The 229 nm excited UVRR, which probes the tyrosine and tryptophan side chain vibrations, shows an increase in the solvent exposure of the tryptophan side chains as the temperature is increased. Our results are consistent with the existence of an intermediate state similar to that calculated by Bursulaya and Brooks [Bursulaya, B. D., and Brooks, C. L. (1999) J. Am. Chem. Soc. 121, 9947-9951] and exclude the previously proposed two-state cooperative folding mechanism. Betanova's structure appears to be molten globule over the 3-82 degrees C temperature range of our study.  相似文献   

19.
The peptide N-Boc-L-Phe-dehydro-Leu-L-Val-OCH3 was synthesized by the usual workup procedure and finally by coupling the N-Boc-L-Phe-dehydro-Leu-OH to valine methyl ester. It was crystallized from its solution in methanol-water mixture at 4 degrees C. The crystals belong to the triclinic space group P1 with a = 5.972(5) A, b = 9.455(6) A, c = 13.101(6) A, alpha = 103.00(4) degrees, beta = 97.14(5) degrees, gamma = 102.86(5) degrees, V = 690.8(8) A, Z = 1, dm = 1.179(5) Mg m-3 and dc = 1.177(5) Mg m-3. The structure was determined by direct methods using SHELXS86. It was refined by block-diagonal least-squares procedure to an R value of 0.060 for 1674 observed reflections. The C alpha 2-C beta 2 distance of 1.323(9) A in dehydro-Leu is an appropriate double bond length. The bond angle C alpha-C beta-C gamma in the dehydro-Leu residue is 129.4(8) degrees. The peptide backbone torsion angles are theta 1 = -168.6(6) degrees, omega 0 = 170.0(6) degrees, phi 1 = -44.5(9) degrees, psi 1 = 134.5(6) degrees, omega 1 = 177.3(6) degrees, phi 2 = 54.5(9) degrees, psi 2 = 31.1(10) degrees, omega 2 = 171.7(6) degrees, phi 3 = 51.9(8) degrees, psi T3 = 139.0(6) degrees, theta T = -175.7(6) degrees. These values show that the backbone adopts a beta-turn II conformation. As a result of beta-turn, an intramolecular hydrogen bond is formed between the oxygen of the ith residue and NH of the (i + 3)th residue at a distance of 3.134(6) A.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Synthesis of the beta-carotene oxidation product, 2,3-dihydro-5,8-endoperoxy-beta-apo-carotene-13-one (1) was achieved in six steps starting from beta-ionone. Photo-oxygenation of all trans-retinoic acid (8) and 13-cis-retinoic acid (9) produced a mixture of 5S*,8S*-epidioxy-5,8-dihydroretinoic acid (10) and 13-cis-5S*,8S*-epidioxy-5,8-dihydroretinoic acid (11). Methylation of the crude photo-oxygenation mixture afforded the corresponding methyl esters 12 and 13, respectively, both of which underwent ready aerial oxidation yielding hitherto unknown oxidation products of retinoic acid identified as methyl 5S*,8S*-epidioxy-9,10beta-epoxy-5,8,9,10-tetrahydroretinoate (14) and methyl 13-cis-5S*,8S*-epidioxy-9,10beta-epoxy-5,8,9,10-tetrahydroretinoate (15). Evaluation of 1, all trans-retinoic acid (8), 13-cis-retinoic acid (9), and the photo-oxygenation products 10-15 in a panel of five cancer cell lines showed 1 to be inactive and that 11 is significantly cytotoxic compared with the other retinoic acid analogs suggesting the requirement of the carboxylic acid moiety and the cis-geometry of the 13(14) double bond for cytotoxic activity.  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