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 30 毫秒
1.
Ito H  Tanaka S  Miyasaka M 《Biopolymers》2002,65(2):61-80
We utilize electrophoresis and find that a thermally treated equimolar mixture of the oligonucleotide d(G(5)T(5)) and its complementary oligonucleotide d(A(5)C(5)) exhibits either two bands or a single band in one lane, depending on the conditions of the incubation solutions. The thermally treated d(G(5)T(5)) solution loaded in a different lane exhibits a single band of the parallel quadruplex [d(G(5)T(5))](4), which is composed of homocyclic hydrogen-bonded G(4) and T(4) tetrads previously proposed. For the thermally treated equimolar mixture of d(G(5)T(5)) and d(A(5)C(5)), the fast band is assigned to a Watson-Crick d(G(5)T(5)). d(A(5)C(5)) duplex, so that the slow band with the same low mobility as that of [d(G(5)T(5))](4) may be assigned to either [d(G(5)T(5))](4) itself or a [d(G(5)T(5)). d(A(5)C(5))](2) quadruplex. If the latter compound is true, this may be the antiparallel quadruplex composed of the heterocyclic hydrogen-bonded G-C-G-C and T-A-T-A tetrads proposed previously. After removing these three bands for the duplex and two kinds of hypothetical quadruplexes, we electrophoretically elute the corresponding compounds in the same electrophoresis buffer using an electroeluter. The eluted compounds are ascertained to be stable by electrophoresis. The circular dichroism (CD) and UV absorption spectra measured for the three isolated compounds are found to be clearly different. For the electrophoretic elution of the hypothetical [d(G(5)T(5))](4) quadruplex, the result of the molecularity of n = 4 obtained from the CD melting curve analysis provides further support for the formation of the parallel [d(G(5)T(5))](4) quadruplex already proposed. For the thermally treated equimolar mixture of d(G(5)T(5)) and d(C(5)A(5)), the fast band with a molecularity of n = 2 corresponds to the Watson-Crick duplex, d(G(5)T(5)). d(A(5)C(5)). The slow band with a molecularity of n = 4 indicates the antiparallel quadruplex [d(G(5)T(5)). d(A(5)C(5))](2), whose observed CD and UV spectra are different from those of [d(G(5)T(5))](4). By electrophoresis, after reannealing the eluted compound [d(G(5)T(5)). d(A(5)C(5))](2), a distinct photograph showing the band splitting of this quadruplex band into the lower duplex and upper quadruplex bands is not possible; but by a transilluminator, we occasionally observe this band splitting with the naked eye. The linear response polarizability tensor calculations for the thus determined structures of the [d(G(5)T(5))](4) quadruplex, the McGavin-like [d(G(5)T(5)). d(A(5)C(5))](2) quadruplex, and the Watson-Crick d(G(5)T(5)). d(A(5)C(5)) duplex are found to qualitatively predict the observed CD and UV spectra.  相似文献   

2.
Incubation of linoleic acid with the 105,000g particle fraction of the homogenate of the broad bean (Vicia faba L.) led to the formation of the following products: 13(S)-hydroxy-9(Z),11(E)-octadecadienoic acid, 9,10-epoxy-12(Z)-octadecenoic acid (9(R),10(S)/9(S)/10(R), 80/20), 12,13-epoxy-9(Z)-octadecenoic acid (12(S),13(R)/12(R)/13(S), 64/36), and 9,10-epoxy-13(S)-hydroxy-11(E)-octadecenoic acid (9(S),10(R)/9(R),10(S), 91/9). Oleic acid incubated with the enzyme preparation in the presence of 13(S)-hydroperoxy-9(Z),11(E)-octadecadienoic acid or cumene hydroperoxide was converted into 9,10-epoxyoctadecanoic acid (9(R),10(S)/9(S),10(R), 79/21). Two enzyme activities were involved in the formation of the products, an omega 6-lipoxygenase and a hydroperoxide-dependent epoxygenase. The lipoxygenase, but not the epoxygenase, was inhibited by low concentrations of 5,8,11,14-eicosatetraynoic acid and nordihydroguaiaretic acid. In contrast, the epoxygenase, but not the lipoxygenase, was readily inactivated in the presence of 13(S)-hydroperoxy-9(Z),11(E)-octadecadienoic acid. Studies with 18O2-labeled 13(S)-hydroperoxy-9(Z),11(E)-octadecadienoic acid showed that the epoxide oxygens of 9,10-epoxyoctadecanoic acid and of 9,10-epoxy-13(S)-hydroxy-11(E)-octadecenoic acid were derived from hydroperoxide and not from molecular oxygen.  相似文献   

3.
Ex vivo ?(13)C, (2)H? NMR spectroscopy allowed to estimate the relative sizes of neuronal and glial glutamate pools and the relative contributions of (1-(13)C) glucose and (2-(13)C, 2-(2)H(3)) acetate to the neuronal and glial tricarboxylic acid cycles of the adult rat brain. Rats were infused during 60 min in the right jugular vein with solutions containing (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose or (2-(13)C, 2-(2)H(3)) acetate only. At the end of the infusion the brains were frozen in situ and perchloric acid extracts were prepared and analyzed by high resolution (13)C NMR spectroscopy (90.5 MHz). The relative sizes of the neuronal and glial glutamate pools and the contributions of acetyl-CoA molecules derived from (2-(13)C, (2)H(3)) acetate or (1-(13)C) glucose entering the tricarboxylic acid cycles of both compartments, could be determined by the analysis of (2)H-(13)C multiplets and (2)H induced isotopic shifts observed in the C4 carbon resonances of glutamate and glutamine. During the infusions with (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose, the glial glutamate pool contributed 9% of total cerebral glutamate being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (4%), (2-(13)C) acetyl-CoA (3%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (2%). The neuronal glutamate pool accounted for 91% of the total cerebral glutamate being mainly originated from (2-(13)C) acetyl-CoA (86%) and (2-(13)C, 2-(2)H) acetyl-CoA (5%). During the infusions of (2-(13)C, 2-(2)H(3)) acetate only, the glial glutamate pool contributed 73% of the cerebral glutamate, being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (36%), (2-(13)C, 2-(2)H) acetyl-CoA (27%) and (2-(13)C) acetyl-CoA (10%). The neuronal pool contributed 27% of cerebral glutamate being formed from (2-(13)C) acetyl-CoA (11%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (16%). These results illustrate the potential of ?(13)C, (2)H? NMR spectroscopy as a novel approach to investigate substrate selection and metabolic compartmentation in the adult mammalian brain.  相似文献   

4.
5.
A novel gene (Cacng2; gamma(2)) encoding a protein similar to the voltage-activated Ca(2+) channel gamma(1) subunit was identified as the defective gene in the epileptic and ataxic mouse, stargazer. In this study, we analyzed the association of this novel neuronal gamma(2) subunit with Ca(2+) channels of rabbit brain, and the function of the gamma(2) subunit in recombinant neuronal Ca(2+) channels expressed in Xenopus oocytes. Our results showed that the gamma(2) subunit and a closely related protein (called gamma(3)) co-sedimented and co-immunoprecipitated with neuronal Ca(2+) channel subunits in vivo. Electrophysiological analyses showed that gamma(2) co-expression caused a significant decrease in the current amplitude of both alpha(1B)(alpha(1)2.2)-class (36.8%) and alpha(1A)(alpha(1)2.1)-class (39.7%) Ca(2+) channels (alpha(1)beta(3)alpha(2)delta). Interestingly, the inhibitory effects of the gamma(2) subunit on current amplitude were dependent on the co-expression of the alpha(2)delta subunit. In addition, co-expression of gamma(2) or gamma(1) also significantly decelerates the activation kinetics of alpha(1B)-class Ca(2+) channels. Taken together, these results suggest that the gamma(2) subunit is an important constituent of the neuronal Ca(2+) channel complex and that it down-regulates neuronal Ca(2+) channel activity. Furthermore, the gamma(2) subunit likely contributes to the fine-tuning of neuronal Ca(2+) channels by counterbalancing the effects of the alpha(2)delta subunit.  相似文献   

6.
King RB 《Chirality》2001,13(8):465-473
Chiral molecules can either be handed (i.e., "shoes") or nonhanded ("potatoes"). The only chiral ligand partition for tetrahedral metal complexes (or for a tetrahedral carbon atom such as that found in amino acids and other chiral biological molecules) is the fully unsymmetrical degree 6 partition (1(4)), which leads to handed metal complexes of the type MABCD with a lowest-degree chirality polynomial consisting of the product of all six possible linear factors of the type (s(i)-s(j)) where 1 < or = i,j < or = 4. The lowest-degree chiral ligand partitions for octahedral metal complexes are the degree 6 partitions (31(3)) and (2(3)) leading to handed chiral metal complexes of the types fac-MA(3)BCD and cis-MA(2)B(2)C(2). The form of the lowest-degree chirality polynomial for the (31(3)) chiral ligand partition of the octahedron resembles that of the (1(4)) chiral ligand partition of the tetrahedron, likewise with four different ligands. However, the form of the lowest-degree chirality polynomial for the (2(3)) chiral ligand partition of the octahedron corresponds to the square of the chirality polynomial of the (1(3)) chiral ligand partition of the polarized triangle, which likewise has three different ligands. Ligand partitions for octahedral metal complexes such as (2(2)1(2)), (21(4)), and (1(6)), which are less symmetrical than the lowest-degree chiral ligand partitions (31(3)) and (2(3)), lead to chiral octahedral metal complexes which are nonhanded. In such complexes, pairs of enantiomers can be interconverted by simple ligand interchanges without ever going through an achiral intermediate.  相似文献   

7.
The reaction of copper(II) complexes supported by a series of beta-diketiminate ligands ((R1,R2)L, [(Dipp)N-C(R(2))-C(R(1))-C(R(2))-N(Dipp)](-), Dipp=2,6-diisopropylphenyl; see ) and H(2)O(2) has been examined spectroscopically at a low temperature. The beta-diketiminatocopper(II) complexes with R(2)=H (no substituent on the beta-carbon) provided a copper-oxygen intermediate that exhibited the same spectroscopic features as those of the bis(mu-oxo)dicopper(III) complex generated by the reaction of corresponding beta-diketiminatocopper(I) complex and O(2). On the other hand, the beta-diketiminatocopper(II) complexes with methyl substituent on the beta-carbon (R(2)=Me) did not produce such an intermediate in the same reaction. The beta-diketiminatocopper(II) complexes carrying an electron-withdrawing substituent on the alpha-carbon (R(1)=NO(2) or CN) but no beta-substituent (R(2)=H) exhibited a high catalytic activity in the oxygenation reaction of alkanes with H(2)O(2). Mechanism of the catalytic oxygenation reaction as well as the substituent effects of the ligands on the copper(II)-H(2)O(2) reactivity is discussed.  相似文献   

8.
The organometallic precursor (NEt(4))(2)[ReBr(3)(CO)(3)] was reacted with bidendate dithioethers (L) of the general formula H(3)C-S-CH(2)CH(2)-S-R (R = -CH(2)CH(2)COOH, CH(2)-C&tbd1;CH) and R'-S-CH(2)CH(2)-S-R' (R' = CH(3)CH(2)-, CH(3)CH(2)-OH, and CH(2)COOH) in methanol to form stable rhenium(I) tricarbonyl complexes of the general composition [ReBr(CO)(3)L]. Under these conditions, the functional groups do not participate in the coordination. As a prototypic representative of this type of Re compounds, the propargylic group bearing complex [ReBr(CO(3))(H(3)C-S-CH(2)CH(2)-S-CH(2)C&tbd1;CH)] Re2 was studied by X-ray diffraction analysis. Its molecular structure exhibits a slightly distorted octahedron with facial coordination of the carbonyl ligands. The potentially tetradentate ligand HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH was reacted with the trinitrato precursor [Re(NO(3))(3)(CO)(3)](2-) to yield a cationic complex [Re(CO)(3)(HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH)]NO(3) Re8 which shows the coordination of one hydroxy group. Re8 has been characterized by correct elemental analysis, infrared spectroscopy, capillary electrophoresis, and X-ray diffraction analysis. Ligand exchange reaction of the carboxylic group bearing ligands H(3)C-S-CH(2)CH(2)-S-CH(2)CH(2)-COOH and HOOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH with (NEt(4))(2)[ReBr(3)(CO)(3)] in water and with equimolar amounts of NaOH led to complexes in which the bromide is replaced by the carboxylic group. The X-ray structure analysis of the complex [Re(CO)(3)(OOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH)] Re6 shows the second carboxylic group noncoordinated offering an ideal site for functionalization or coupling a biomolecule. The no-carrier-added preparation of the analogous (99m)Tc(I) carbonyl thioether complexes could be performed using the precursor fac-[(99m)Tc(H(2)O)(3)(CO)(3)](+), with yields up to 90%. The behavior of the chlorine containing (99m)Tc complex [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 in aqueous solution at physiological pH value was investigated. In saline, the chromatographically separated compound was stable for at least 120 min. However, in chloride-free aqueous solution, a water-coordinated cationic species Tc1a of the proposed composition [(99m)Tc(H(2)O)(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))](+) occurred. The cationic charge of the conversion product was confirmed by capillary electrophoresis. By the introduction of a carboxylic group into the thioether ligand as a third donor group, the conversion could be suppressed and thus the neutrality of the complex preserved. Biodistribution studies in the rat demonstrated for the neutral complexes [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 and [(99m)TcCl(CO)(3)(CH(2)-S-CH(2)CH(2)-S-CH(2)-C&tbd1;CH)] Tc2 a significant initial brain uptake (1.03 +/- 0.25% and 0.78 +/- 0.08% ID/organ at 5 min. p.i.). Challenge experiments with glutathione clearly indicated that no transchelation reaction occurs in vivo.  相似文献   

9.
We report the results of NMR studies and computer simulations of potent antagonists reflective of the alpha(IIb)beta(3) receptor-bound conformations. The peptides c[Mpa-(15)N-Arg(1)-(15)N-Gly(2)-(15)N-Asp(3)-(15)N-Phe(4)-(15)N-Arg(5)-Cys]-NH(2) (Phe-Arg analog) (Mpa: 3-mercaptopropionic acid) and c[Mpa-(15)N-Arg(1)-(15)N-Gly(2)-(15)N-Asp(3)-(15)N-Asp(4)-(15)N-Val(5)-Cys]-NH(2) (Asp-Val analog) were subjected to (15)N-edited NMR experiments to study the conformations of these peptides in the absence and in the presence of alpha(IIb)beta(3) receptor. The NMR studies of the Phe-Arg analog, a selective alpha(IIb)beta(3) antagonist, resulted in distinctly different experimental data in the presence and absence of the receptor. The computer simulations for this peptide resulted in one large family of structures consistent with the experimental data. This conformation suggests a type I beta-turn spanning residues Arg(1) and Gly(2) when bound to the receptor and we were able to establish a model for the three dimensional arrangement of the pharmacophores. The studies on the Asp-Val analog, an alpha(v)beta(3) antagonist that binds to the alpha(IIb)beta(3) with moderate affinity, resulted in conformations that are not as well defined as those for the Phe-Arg analog but are consistent with the model established for this analog. These results are important for the design of novel alpha(IIb)beta(3) antagonists.  相似文献   

10.
The catalytic activity of the bacterial phosphotriesterase (PTE) toward a series of chiral analogues of the chemical warfare agents sarin and soman was measured. Chemical procedures were developed for the chiral syntheses of the S(P)- and R(P)-enantiomers of O-isopropyl p-nitrophenyl methylphosphonate (sarin analogue) in high enantiomeric excess. The R(P)-enantiomer of the sarin analogue (k(cat)=2600 s(-1)) was the preferred substrate for the wild-type PTE relative to the corresponding S(P)-enantiomer (k(cat)=290 s(-1)). The observed stereoselectivity was reversed using the PTE mutant, I106A/F132A/H254Y where the k(cat) values for the R(P)- and S(P)-enantiomers were 410 and 4200 s(-1), respectively. A chemo-enzymatic procedure was developed for the chiral synthesis of the four stereoisomers of O-pinacolyl p-nitrophenyl methylphosphonate (soman analogue) with high diastereomeric excess. The R(P)R(C)-stereoisomer of the soman analogue was the preferred substrate for PTE. The k(cat) values for the soman analogues were measured as follows: R(P)R(C,) 48 s(-1); R(P)S(C), 4.8 s(-1); S(P)R(C), 0.3 s(-1), and S(P)S(C), 0.04 s(-1). With the I106A/F132A/H254Y mutant of PTE the stereoselectivity toward the chiral phosphorus center was reversed. With the triple mutant the k(cat) values for the soman analogues were found to be as follows: R(P)R(C,) 0.3 s(-1); R(P)S(C), 0.3 s(-1); S(P)R(C), 11s(-1), and S(P)S(C), 2.1 s(-1). Prior investigations have demonstrated that the S(P)-enantiomers of sarin and soman are significantly more toxic than the R(P)-enantiomers. This investigation has demonstrated that mutants of the wild-type PTE can be readily constructed with enhanced catalytic activities toward the most toxic stereoisomers of sarin and soman.  相似文献   

11.
Aspergillus parasiticus produces the minor aflatoxins M(1) (AFM(1)), M(2) (AFM(2)), GM(1) (AFGM(1)), and GM(2) (AFGM(2)), as well as the major aflatoxins B(1) (AFB(1)), B(2) (AFB(2)), G(1) (AFG(1)), and G(2) (AFG(2)). Feeding of A. parasiticus with aspertoxin (12c-hydroxyOMST) caused AFM(1) and AFGM(1), and cell-free experiments using the microsomal fraction of A. parasiticus and aspertoxin caused production of AFM(1), indicating that aspertoxin is a precursor of AFM(1) and AFGM(1). Feeding of the same fungus with O-methylsterigmatocystin (OMST) caused AFM(1) and AFGM(1) together with AFB(1) and AFG(1); feeding with dihydroOMST (DHOMST) caused AFM(2) and AFGM(2) together with AFB(2) and AFG(2). Incubation of either the microsomal fraction or OrdA enzyme-expressing yeast with OMST caused production of aspertoxin together with AFM(1) and AFB(1). These results demonstrated that the OrdA enzyme catalyzes both 12c-hydroxylation reaction from OMST to aspertoxin and the successive reaction from aspertoxin to AFM(1). In contrast, feeding of the fungus with AFB(1) did not produce any AFM(1), demonstrating that M-/GM-aflatoxins are not produced from B-/G-aflatoxins. Furthermore, AFM(1) together with AFB(1) and AFG(1) was also produced from 11-hydroxyOMST (HOMST) in feeding experiment of A. parasiticus, whereas no aflatoxins were produced when used the ordA deletion mutant. These results demonstrated that OrdA enzyme can also catalyze 12c-hydroxylation of HOMST to produce 11-hydroxyaspertoxin, which serves as a precursor for the production of AFM(1) and AFGM(1). The same pathway may work for the production of AFM(2) and AFGM(2) from DHOMST and dihydroHOMST through the formation of dihydroaspertoxin and dihydro-11-hydroxyaspertoxin, respectively.  相似文献   

12.
YddV from Escherichia coli (Ec) is a novel globin-coupled heme-based oxygen sensor protein displaying diguanylate cyclase activity in response to oxygen availability. In this study, we quantified the turnover numbers of the active [Fe(III), 0.066 min(-1); Fe(II)-O(2) and Fe(II)-CO, 0.022 min(-1)] [Fe(III), Fe(III)-protoporphyrin IX complex; Fe(II), Fe(II)-protoporphyrin IX complex] and inactive forms [Fe(II) and Fe(II)-NO, <0.01 min(-1)] of YddV for the first time. Our data indicate that the YddV reaction is the rate-determining step for two consecutive reactions coupled with phosphodiesterase Ec DOS activity on cyclic di-GMP (c-di-GMP) [turnover number of Ec DOS-Fe(II)-O(2), 61 min(-1)]. Thus, O(2) binding and the heme redox switch of YddV appear to be critical factors in the regulation of c-di-GMP homeostasis. The redox potential and autoxidation rate of heme of the isolated heme domain of YddV (YddV-heme) were determined to be -17 mV versus the standard hydrogen electrode and 0.0076 min(-1), respectively. The Fe(II) complexes of Y43A and Y43L mutant proteins (residues at the heme distal side of the isolated heme-bound globin domain of YddV) exhibited very low O(2) affinities, and thus, their Fe(II)-O(2) complexes were not detected on the spectra. The O(2) dissociation rate constant of the Y43W protein was >150 s(-1), which is significantly larger than that of the wild-type protein (22 s(-1)). The autoxidation rate constants of the Y43F and Y43W mutant proteins were 0.069 and 0.12 min(-1), respectively, which are also markedly higher than that of the wild-type protein. The resonance Raman frequencies representing ν(Fe-O(2)) (559 cm(-1)) of the Fe(II)-O(2) complex and ν(Fe-CO) (505 cm(-1)) of the Fe(II)-CO complex of Y43F differed from those (ν(Fe-O(2)), 565 cm(-1); ν(Fe-CO), 495 cm(-1)) of the wild-type protein, suggesting that Tyr43 forms hydrogen bonds with both O(2) and CO molecules. On the basis of the results, we suggest that Tyr43 located at the heme distal side is important for the O(2) recognition and stability of the Fe(II)-O(2) complex, because the hydroxyl group of the residue appears to interact electrostatically with the O(2) molecule bound to the Fe(II) complex in YddV. Our findings clearly support a role of Tyr in oxygen sensing, and thus modulation of overall conversion from GTP to pGpG via c-di-GMP catalyzed by YddV and Ec DOS, which may be applicable to other globin-coupled oxygen sensor enzymes.  相似文献   

13.
Six-coordinate low-spin iron(III) porphyrinates adopt either common (d(xy))(2)(d(xz),d(yz))(3) or less common (d(xz),d(yz))(4)(d(xy))(1) ground state. In this review article, three major factors that affect the electronic ground state have been examined. They are (i) nature of the axial ligand, (ii) electronic effect of peripheral substituents, and (iii) deformation of porphyrin ring. On the basis of the (1)H NMR, (13)C NMR, and EPR data, it is now clear that (i) the axial ligands with low-lying pi* orbitals such as tert-butylisocyanide and 4-cyanopyridine, (ii) the electron donating groups at the meso-carbon atoms, and (iii) the ruffled deformation of porphyrin ring stabilize the (d(xz),d(yz))(4)(d(xy))(1) ground state. By manipulating these factors, we are able to prepare various low-spin iron(III) porphyrinates with unusual electronic structures such as bis(imidazole) complexes with the (d(xz),d(yz))(4)(d(xy))(1) ground state or bis(tert-butylisocyanide) complexes with the (d(xy))(2)(d(xz),d(yz))(3) ground state; bis(imidazole) and bis(tert-butylisocyanide) complexes usually adopt the (d(xy))(2)(d(xz),d(yz))(3) and (d(xz),d(yz))(4)(d(xy))(1) ground state, respectively.  相似文献   

14.
The N-terminal 1-34 fragments of the parathyroid hormone (PTH) and parathyroid hormone-related protein (PTHrP) elicit the full spectrum of bone-related biological activities of the intact native sequences. It has been suggested that the structural elements essential for bioactivity are two helical segments located at the N-terminal and C-terminal sequences, connected by hinges or flexible points around positions 12 and 19. In order to assess the relevance of the local conformation around Gly(12) upon biological function, we synthesized and characterized the following PTH(1-34) analogues containing Aib residues: (I) A-V-S-E-I-Q-F-nL-H-N-Aib-G-K-H-L-S-S-nL-E-R-V-E-Nal-L-R-K-K-L-Q-D-V-H-N-Y-NH(2) ([Nle(8,18), Aib(11), Nal(23),Tyr(34)]bPTH(1-34)-NH(2)); (II) A-V-S-E-I-Q-F-nL-H-N-L-Aib-K-H-L-S-S-nL-E-R-V-E-Nal-L-R-K-K-L-Q-D-V-H-N-Y-NH(2) ([Nle(8,18), Aib(12),Nal(23),Tyr(34)]bPTH(1-34)-NH(2)); (III) A-V-S-E-I-Q-F-nL-H-N-L-G-Aib-H-L-S-S-nL-E-R-V-E-Nal-L-R-K-K-L-Q-D-V-H-N-Y-NH(2) ([Nle(8,18), Aib(13), Nal(23),Tyr(34)]bPTH(1-34)-NH(2)); (IV) A-V-S-E-I-Q-F-nL-H-N-Aib-Aib-K-H-L-S-S-nL-E-R-V-E-Nal-L-R-K-K-L-Q-D-V-H-N-YNH(2) ([Nle(8,18), Aib(11,12), Nal(23),Tyr(34)]bPTH(1-34)-NH(2)); (V) A-V-S-E-I-Q-F-nL-H-N-L-Aib-Aib-H-L-S-S-nL-E-R-V-E-Nal-L-R-K-K-L-Q-D-V-H-N-Y-NH(2) ([Nle(8,18), Aib(12,13),Nal(23),Tyr(34)]bPTH(1-34)-NH(2)). (nL= Nle; Nal= L-(2-naphthyl)-alanine; Aib= alpha-amino-isobutyric acid.) The introduction of Aib residues at position 11 in analogue I or at positions 11 and 12 in analogue IV resulted in a 5-20-fold lower efficacy and a substantial loss of binding affinity compared to the parent compound [Nle(8,18), Nal(23),Tyr(34)]bPTH(1-34)-NH(2). Both binding affinity and adenylyl cyclase stimulation activity are largely restored when the Aib residues are introduced at position 12 in analogue II, 13 in analogue III, and 12-13 in analogue V. The conformational properties of the analogues in aqueous solution containing dodecylphosphocholine micelles were studied by CD, two-dimensional (2D) NMR and computer simulations. The results indicated the presence of two helical segments in all analogues, located at the N-terminal and C-terminal sequences. Insertion of Aib residues at positions 12 and 13, or of Aib dyads at positions 11-12 and 12-13, enhances the stability of the N-terminal helix of all analogues. In all analogues the Aib residues are included in the helical segments. These results confirmed the importance of the helical structure in the N-terminal activation domain, as well as of the presence of the Leu(11) hydrophobic side chain in the native sequence, for PTH-like bioactivity.  相似文献   

15.
NMR was used to study the solution structure of bovine tRNA(Trp) hyperexpressed in Escherichia coli. With the use of (15)N labeling and site-directed mutagenesis to assign overlapping resonances through the base pair replacement of U(71)A(2) by G(2)C(71), U(27)A(43) by G(27)C(43), and G(12)C(23) by U(12)A(23), the resonances of all 26 observable imino protons in the helical regions and in the tertiary interactions were assigned unambiguously by means of two-dimensional nuclear Overhauser effect spectroscopy and heteronuclear single quantum coherence methods. When the discriminator base A(73) and the G(12)C(23) base pair on the D stem, two identity elements on bovine tRNA(Trp) that are important for effective recognition by tryptophanyl-tRNA synthetase, were mutated to the ineffective forms of G(73) and U(12)A(23), respectively, NMR analysis revealed an important conformational change in the U(12)A(23) mutant but not in the G(73) mutant molecule. Thus A(73) appears to be directly recognized by tryptophanyl-tRNA synthetase, and G(12)C(23) represents an important structural determinant. Mg(2+) effects on the assigned resonances of imino protons allowed the identification of strong, medium, and weak Mg(2+) binding sites in tRNA(Trp). Strong Mg(2+) binding modes were associated with the residues G(7), s(4)U(8) (where s(4)U is 4-thiouridine), G(12), and U(52). The observations that G(42) was associated with strong Mg(2+) binding in only the U(12)A(23) mutant tRNA(Trp) but not the wild type or G(73) mutant tRNA(Trp) and that the G(7), s(4)U(8), G(24), and G(22) imino protons are associated with a two-site Mg(2+) binding mode in wild type and G(73) mutant but only a one-site mode in the U(12)A(23) mutant established the occurrence of conformational change in the U(12)A(23) mutant tRNA(Trp). These observations also established the dependence of Mg(2+) binding on tRNA conformation and the usefulness of Mg(2+) binding sites as conformational probes. The thermal titration of tRNA(Trp) in the presence and absence of 10 mm Mg(2+) indicated that overall tRNA(Trp) structure stability was increased by more than 15 degrees C by the presence of Mg(2+).  相似文献   

16.
Chu HA  Feng YW  Wang CM  Chiang KA  Ke SC 《Biochemistry》2004,43(34):10877-10885
Light-induced Fourier transform infrared difference spectroscopy has been applied to studies of ammonia effects on the oxygen-evolving complex (OEC) of photosystem II (PSII). We found that NH(3) induced characteristic spectral changes in the region of the symmetric carboxylate stretching modes (1450-1300 cm(-1)) of the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectra of PSII. The S(2) state carboxylate mode at 1365 cm(-1) in the S(2)Q(A)(-)/S(1)Q(A) spectrum of the controlled samples was very likely upshifted to 1379 cm(-1) in that of NH(3)-treated samples; however, the frequency of the corresponding S(1) carboxylate mode at 1402 cm(-1) in the same spectrum was not significantly affected. These two carboxylate modes have been assigned to a Mn-ligating carboxylate whose coordination mode changes from bridging or chelating to unidentate ligation during the S(1) to S(2) transition [Noguchi, T., Ono, T., and Inoue, Y. (1995) Biochim. Biophys. Acta 1228, 189-200; Kimura, Y., and Ono, T.-A. (2001) Biochemistry 40, 14061-14068]. Therefore, our results show that NH(3) induced significant structural changes of the OEC in the S(2) state. In addition, our results also indicated that the NH(3)-induced spectral changes of the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are dependent on the temperature of the FTIR measurement. Among the temperatures we measured, the strongest effect was seen at 250 K, a lesser effect was seen at 225 K, and little or no effect was seen at 200 K. Furthermore, our results also showed that the NH(3) effects on the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are dependent on the concentrations of NH(4)Cl. The NH(3)-induced upshift of the 1365 cm(-1) mode is apparent at 5 mM NH(4)Cl and is completely saturated at 100 mM NH(4)Cl concentration. Finally, we found that CH(3)NH(2) has a small but clear effect on the spectral change of the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectrum of PSII. The effects of amines on the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectra (NH(3) > CH(3)NH(2) > AEPD and Tris) are inverse proportional to their size (Tris approximately AEPD > CH(3)NH(2) > NH(3)). Therefore, our results showed that the effects of amines on the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are sterically selective for small amines. On the basis of the correlations between the conditions (dependences on the excitation temperature and NH(3) concentration and the steric requirement for the amine effects) that give rise to the NH(3)-induced upshift of the 1365 cm(-)(1) mode in the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII and the conditions that give rise to the altered S(2) state multiline EPR signal, we propose that the NH(3)-induced upshift of the 1365 cm(-1) mode is caused by the binding of NH(3) to the site on the Mn cluster that gives rise to the altered S(2) state multiline EPR signal. In addition, we found no significant NH(3)-induced change in the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectrum at 200 K. Under this condition, the OEC gives rise to the NH(3)-stabilized g = 4.1 EPR signal and a suppressed g = 2 multiline EPR signal. Our results suggest that the structural difference of the OEC between the normal g = 2 multiline form and the NH(3)-stabilized g = 4.1 form is small.  相似文献   

17.
This minireview article highlights the energetics and the dynamics of the 1(1)B(u)(-) and 3(1)A(g)(-) states of carotenoids discovered very recently. Those "hidden" covalent states have been revealed by measurements of resonance-Raman excitation profiles of crystalline carotenoids. The dependence of the energies of the low-lying singlet states, including the 1(1)B(u)(+), 3(1)A(g)(-), 1(1)B(u)(-), and 2(1)A(g)(-) states, on the number of conjugated double bonds (n) is in agreement with the extrapolation of those state energies calculated by Tavan and Schulten for shorter polyenes (P. Tavan and K. Schulten, Journal of Chemical Physics, 1986, vol. 85, pp. 6602-6609). It has also been shown that the internal-conversion processes among those singlet states take place in accord with the state ordering, i.e., 1(1)B(u)(+) --> 1(1)B(u)(-) --> 2(1)A(g)(-) --> 1(1)A(g)(-) (the ground state) for carotenoids having n = 9 and 10, whereas 1(1)B(u)(+) --> 3(1)A(g)(-) --> 1(1)B(u) (-) --> 2(1)A(g)(-) --> 1(1)A(g)(-) for carotenoids having n = 11-13. Radiative transitions of 1(1)B(u)(+) --> 2(1)A(g)(-) and 1(1)B(u)(-) --> 2(1)A(g)(-) as well as a branching into the triplet manifold of 1(1)B(u)(-) --> 1(3)A(g) --> 1(3)B(u) have also been found. Those low-lying singlet states of all-trans carotenoids can facilitate multiple channels of singlet-energy transfer to bacteriochlorophyll in the LH2 antenna complexes of purple photosynthetic bacteria. Thus, the newly found 1(1)B(u)(-) and 3(1)A(g)(-) states of carotenoids need to be incorporated into the picture of carotenoid-to-bacteriochlorophyll singlet-energy transfer.  相似文献   

18.
Phosphoinositide (3,5)-bisphosphate [PI(3,5)P(2)] is a newly identified phosphoinositide that modulates intracellular Ca(2+) by activating ryanodine receptors (RyRs). Since the contractile state of arterial smooth muscle depends on the concentration of intracellular Ca(2+), we hypothesized that by mobilizing sarcoplasmic reticulum (SR) Ca(2+) stores PI(3,5)P(2) would increase intracellular Ca(2+) in arterial smooth muscle cells and cause vasocontraction. Using immunohistochemistry, we found that PI(3,5)P(2) was present in the mouse aorta and that exogenously applied PI(3,5)P(2) readily entered aortic smooth muscle cells. In isolated aortic smooth muscle cells, exogenous PI(3,5)P(2) elevated intracellular Ca(2+), and it also contracted aortic rings. Both the rise in intracellular Ca(2+) and the contraction caused by PI(3,5)P(2) were prevented by antagonizing RyRs, while the majority of the PI(3,5)P(2) response was intact after blockade of inositol (1,4,5)-trisphosphate receptors. Depletion of SR Ca(2+) stores with thapsigargin or caffeine and/or ryanodine blunted the Ca(2+) response and greatly attenuated the contraction elicited by PI(3,5)P(2). The removal of extracellular Ca(2+) or addition of verapamil to inhibit voltage-dependent Ca(2+) channels reduced but did not eliminate the Ca(2+) or contractile responses to PI(3,5)P(2). We also found that PI(3,5)P(2) depolarized aortic smooth muscle cells and that LaCl(3) inhibited those aspects of the PI(3,5)P(2) response attributable to extracellular Ca(2+). Thus, full and sustained aortic contractions to PI(3,5)P(2) required the release of SR Ca(2+), probably via the activation of RyR, and also extracellular Ca(2+) entry via voltage-dependent Ca(2+) channels.  相似文献   

19.
López-Fanjul C  Fernández A  Toro MA 《Genetics》2003,164(4):1627-1633
For neutral additive genes, the quantitative index of population divergence (Q(ST)) is equivalent to Wright's fixation index (F(ST)). Thus, divergent or convergent selection is usually invoked, respectively, as a cause of the observed increase (Q(ST) > F(ST)) or decrease (Q(ST) < F(ST)) of Q(ST) from its neutral expectation (Q(ST) = F(ST)). However, neutral nonadditive gene action can mimic the additive expectations under selection. We have studied theoretically the effect of consecutive population bottlenecks on the difference F(ST) - Q(ST) for two neutral biallelic epistatic loci, covering all types of marginal gene action. With simple dominance, Q(ST) < F(ST) for only low to moderate frequencies of the recessive alleles; otherwise, Q(ST) > F(ST). Additional epistasis extends the condition Q(ST) < F(ST) to a broader range of frequencies. Irrespective of the type of nonadditive action, Q(ST) < F(ST) generally implies an increase of both the within-line additive variance after bottlenecks over its ancestral value (V(A)) and the between-line variance over its additive expectation (2F(ST)V(A)). Thus, both the redistribution of the genetic variance after bottlenecks and the F(ST) - Q(ST) value are governed largely by the marginal properties of single loci. The results indicate that the use of the F(ST) - Q(ST) criterion to investigate the relative importance of drift and selection in population differentiation should be restricted to pure additive traits.  相似文献   

20.
Vacuolar-type rotary H(+)-ATPase/synthase (V(o)V(1)) from Thermus thermophilus, composed of nine subunits, A, B, D, F, C, E, G, I, and L, has been reconstituted from individually isolated V(1) (A(3)B(3)D(1)F(1)) and V(o) (C(1)E(2)G(2)I(1)L(12)) subcomplexes in vitro. A(3)B(3)D and A(3)B(3) also reconstituted with V(o), resulting in a holoenzyme-like complexes. However, A(3)B(3)D-V(o) and A(3)B(3)-V(o) did not show ATP synthesis and dicyclohexylcarbodiimide-sensitive ATPase activity. The reconstitution process was monitored in real time by fluorescence resonance energy transfer (FRET) between an acceptor dye attached to subunit F or D in V(1) or A(3)B(3)D and a donor dye attached to subunit C in V(o). The estimated dissociation constants K(d) for V(o)V(1) and A(3)B(3)D-V(o) were ~0.3 and ~1 nm at 25 °C, respectively. These results suggest that the A(3)B(3) domain tightly associated with the two EG peripheral stalks of V(o), even in the absence of the central shaft subunits. In addition, F subunit is essential for coupling of ATP hydrolysis and proton translocation and has a key role in the stability of whole complex. However, the contribution of the F subunit to the association of A(3)B(3) with V(o) is much lower than that of the EG peripheral stalks.  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