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.
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.
3.
C(4)-type photosynthesis is known to vary with growth and measurement temperatures. In an attempt to quantify its variability with measurement temperature, the photosynthetic parameters - the maximum catalytic rate of the enzyme ribulose 1.5-bisphosphate carboxylase/oxygenase (Rubisco) (V(cmax)), the maximum catalytic rate of the enzyme phosphoenolpyruvate carboxylase (PEPC) (V(pmax)) and the maximum electron transport rate (J(max)) - were examined. Maize plants were grown in climatic-controlled phytotrons, and the curves of net photosynthesis (A(n)) versus intercellular air space CO(2) concentrations (C(i)), and A(n) versus photosynthetic photon flux density (PPFD) were determined over a temperature range of 15-40 degrees C. Values of V(cmax), V(pmax) and J(max) were computed by inversion of the von Caemmerer & Furbank photosynthesis model. Values of V(pmax) and J(max) obtained at 25 degrees C conform to values found in the literature. Parameters for an Arrhenius equation that best fits the calculated values of V(cmax), V(pmax) and J(max) are then proposed. These parameters should be further tested with C(4) plants for validation. Other model key parameters such as the mesophyll cell conductance to CO(2) (g(i)), the bundle sheath cells conductance to CO(2) (g(bs)) and Michaelis-Menten constants for CO(2) and O(2) (K(c), K(p) and K(o)) also vary with temperature and should be better parameterized.  相似文献   

4.
Capacity and electric resistance of lipid membranes composed of lecithin and cholesterol were determined. The components were chosen for the study because they were present in biological membranes. Capacitance of the lecithin and cholesterol membranes amounts to 0.38 and 0.61 microF/cm(2), and resistance to 1.44(10(4)and 2.12(10(6)Omega cm(2), respectively. A 1:1 complex appears as a result of lecithin-cholesterol membrane formation. Parameters of the membrane formed of the lecithin-cholesterol complex were determined: surface concentration (Gamma(3)), capacitance (C(3)), and conductance (R;(3)(-1), as well as the stability constant (K) of the complex. The mean values of those magnitudes are as follows: 4.265(10(-6)mol/m(2), 0.54 microF/cm(2), 1.381(10(-6)Omega(-1)cm(-2)and 3.748(10(7), respectively.  相似文献   

5.
We report the characterization by resonance Raman spectroscopy of the oxygenated complex (Fe(II)O(2)) of nitric-oxide synthases of Staphylococcus aureus (saNOS) and Bacillus subtilis (bsNOS) saturated with N(omega)-hydroxy-l-arginine. The frequencies of the nu(Fe-O) and nu(O-O) modes were 530 and 1135 cm(-), respectively, in both the presence and absence of tetrahydrobiopterin. On the basis of a comparison of these frequencies with those of saNOS and bsNOS saturated with l-arginine (nu(Fe-O) at 517 cm(-1) and nu(O-O) at 1123 cm(-1)) and those of substrate-free saNOS (nu(Fe-O) at 517 and nu(O-O) at 1135 cm(-1)) (Chartier, F. J. M., Blais, S. P., and Couture, M. (2006) J. Biol. Chem. 281, 9953-9962), we propose two models that account for the frequency shift of nu(Fe-O) (but not nu(O-O)) upon N(omega)-hydroxy-l-arginine binding as well as the frequency shift of nu(O-O) (but not nu(Fe-O)) upon l-arginine binding. The implications of these substrate-specific interactions with respect to catalysis by NOSs are discussed.  相似文献   

6.
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   

7.
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   

8.
We have measured the transition temperatures, T(M), and van't Hoff enthalpies, DeltaH(M), of the thermally induced native-to-unfolded (N-to-U) and molten globule-to-unfolded (MG-to-U) transitions of cytochrome c at pressures between 50 and 2200 bar. We have used the pressure dependence of T(M) to evaluate the changes in volume, Delta(v), accompanying each protein transition event as a function of temperature and pressure. From analysis of the temperature and pressure dependences of Delta(v), we have additionally calculated the changes in expansibility, Delta(e), and isothermal compressibility, Delta(k)(T), associated with the thermally induced conformational transitions of cytochrome c. Specifically, if extrapolated to 25 degrees C, the native-to-unfolded (N-to-U) transition is accompanied by changes in volume, Delta(v), expansibility, Delta(e), and isothermal compressibility, Delta(k)(T), of -(5 +/- 3) x 10(-3) cm(3) g(-1), (1.8 +/- 0.3) x 10(-4) cm(3) g(-1) K(-1), and approximately 0 cm(3) g(-1) bar(-1), respectively. The molten globule-to-unfolded (MG-to-U) transition is accompanied by changes in volume, Delta(v), and isothermal compressibility, Delta(k)(T), of -(2.9 +/- 0.3) x 10(-3) cm(3) g(-1) at 40 degrees C and -(1.9 +/- 0.3) x 10(-6) cm(3) g(-1) bar(-1) at 35 degrees C, respectively. By comparing the volumetric properties of the N-to-U and N-to-MG transitions of cytochrome c, we have estimated the properties of the native-to-molten globule (N-to-MG) transition. For the latter transition, the changes in volume, Delta(v), and isothermal compressibility, Delta(k)(T), are approximately 0 cm(3) g(-1) at 40 degrees C and 1.9 cm(3) g(-1) bar(-1) at 35 degrees C, respectively. Our estimate for the change in expansibility, Delta(e), upon the N-to-MG is negative and equal to -(5 +/- 3) x 10(-4) cm(3) g(-1) K(-1). This finding contrasts with the results of previous studies all of which report positive changes in expansibility associated with protein denaturation. In general, our volumetric data permit us to assess the combined effect of temperature and pressure on the stability of various conformational states of cytochrome c.  相似文献   

9.
We have shown previously that norepinephrine (NE) microdialyzed into the preoptic area (POA) of conscious guinea pigs stimulates local PGE(2) release. To identify the cyclooxygenase (COX) isozyme that catalyzes the production of this PGE(2) and the adrenoceptor (AR) subtype that mediates this effect, we microdialyzed for 6 h NE, cirazoline (alpha(1)-AR agonist), and clonidine (alpha(2)-AR agonist) into the POA of conscious guinea pigs pretreated intrapreoptically (intra-POA) with SC-560 (COX-1 inhibitor) or nimesulide or MK-0663 (COX-2 inhibitors) and measured the animals' core temperature (T(c)) and intra-POA PGE(2) responses. Cirazoline induced T(c) rises promptly after the onset of its dialysis without altering PGE(2) levels. NE and clonidine caused early falls followed by late rises of T(c); intra-POA PGE(2) levels were closely correlated with this thermal course. COX-1 inhibition attenuated the clonidine-induced T(c) and PGE(2) falls but not the NE-elicited hyperthermia, but COX-2 inhibition suppressed both the clonidine- and NE-induced T(c) and PGE(2) rises. Coinfused cirazoline and clonidine reproduced the late T(c) rise of clonidine but not its early fall and also not the early rise produced by cirazoline; on the other hand, the PGE(2) responses were similar to those to NE. Prazosin (alpha(1)-AR antagonist) and yohimbine (alpha(2)-AR antagonist) blocked the effects of their respective agonists. These results indicate that alpha(1)- and alpha(2)-AR agonists microdialyzed into the POA of conscious guinea pigs evoke distinct T(c) responses: alpha(1)-AR activation produces quick, PGE(2)-independent T(c) rises, and alpha(2)-AR stimulation causes an early T(c) fall and a late, COX-2/PGE(2)-dependent T(c) rise.  相似文献   

10.
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   

11.
The cooperative O(2)-binding of hemoglobin (Hb) have been assumed to correlate to change in the quaternary structures of Hb: T(deoxy)- and R(oxy)-quaternary structures, having low and high O(2)-affinities, respectively. Heterotropic allosteric effectors have been shown to interact not only with deoxy- but also oxy-Hbs causing significant reduction in their O(2)-affinities and the modulation of cooperativity. In the presence of two potent effectors, L35 and inositol hexaphosphate (IHP) at pH 6.6, Hb exhibits extremely low O(2)-affinities (K(T)=0.0085mmHg(-1) and K(R)=0.011mmHg(-1)) and thus a very low cooperativity (K(R)/K(T)=1.3 and L(0)=2.4). (1)H-NMR spectra of human adult Hb with these two effectors were examined in order to determine the quaternary state of Hb in solution and to clarify the correlation between the O(2)-affinities and the structural change of Hb caused by the heterotropic effectors. At pH 6.9, (1)H-NMR spectrum of deoxy-Hb in the presence of L35 and IHP showed a marker of the T-quaternary structure (the T-marker) at 14ppm, originated from inter- dimeric α(1)β(2)- (or α(2)β(1)-) hydrogen-bonds, and hyperfine-shifted (hfs) signals around 15-25ppm, caused by high-spin heme-Fe(II)s. Upon addition of O(2), the hfs signals disappeared, reflecting that the heme-Fe(II)s are ligated with O(2), but the T-marker signals still remained, although slightly shifted and broadened, under the partial pressure of O(2) (P(O2)) of 760mmHg. These NMR results accompanying with visible absorption spectroscopy and visible resonance Raman spectroscopy reveal that oxy-Hb in the presence of L35 and IHP below pH 7 takes the ligated T-quaternary structure under the P(O2) of 760mmHg. The L35-concentration dependence of the T-marker in the presence of IHP indicates that there are more than one kind of L35-binding sites in the ligated T-quaternary structure. The stronger binding sites are probably intra-dimeric binding sites between α(1)G- and β(1)G-helices, and the other weaker binding site causes the R→T transition without release of O(2). The fluctuation of the tertiary structure of Hb seems to be caused by both the structural perturbation of α(1)β(1) (or α(2)β(2)) intra-dimeric interface, where the stronger L35-binding sites exist, and by the IHP-binding to the α(1)α(2)- (or β(1)β(2)-) cavity. The tertiary structural fluctuation induced by the allosteric effectors may contribute to the significant reduction of the O(2)-affinity of oxy-Hb, which little depends on the quaternary structures. Therefore, the widely held assumptions of the structure-function correlation of Hb - [the deoxy-state]=[the T-quaternary structure]=[the low O(2)-affinity state] and [the oxy-state]=[the R-quaternary structure]=[the high O(2)-affinity state] and the O(2)-affiny of Hb being regulated by the T/R-quaternary structural transition - are no longer sustainable. This article is part of a Special Issue entitled: Allosteric cooperativity in respiratory proteins.  相似文献   

12.
The corticotropin-releasing factor (CRF) type 1 receptors (CRF(1)) from human (hCRF(1)) and Xenopus (xCRF(1)) differ from one another by their agonist- and antagonist-binding preference. While the agonist-binding site of the xCRF(1) receptor has been mapped, the amino acids that mediate binding of the potent peptide antagonist astressin are unknown. By constructing receptor chimeras followed by site-directed mutagenesis, the astressin-binding site of the xCRF(1) receptor was located between residues 76 and 83. This region partially overlaps with the agonist-selective domain of the xCRF(1) receptor (residues 76-89). Mutagenesis of the amphibian residues Gln(76), Gly(81) and Val(83) to the human sequence (Arg(76)Asn(81)Gly(83)) generated a receptor mutant that bound astressin with even higher affinity than the native hCRF(1) receptor. An amino acid doublet (Glu(70)Tyr(71)) that is conserved in the xCRF(1) and hCRF(2(a)) receptor after incorporation into the hCRF(1) receptor sequence was found to facilitate antagonist binding up to 15-fold higher. In agreement with the binding data, astressin was a more potent functional antagonist at receptors expressing the Glu(70)Tyr(71) motif. These data show that the agonist- and antagonist-binding sites of the hCRF(1) receptor partially overlap and that two amino acids within the N terminus of the hCRF(1) receptor negatively influence binding and functional antagonism of astressin.  相似文献   

13.
Energy-coupled transporters in the outer membrane of Escherichia coli and other Gram-negative bacteria allow the entry of scarce substrates, toxic proteins, and bacterial viruses (phages) into the cells. The required energy is derived from the proton-motive force of the cytoplasmic membrane, which is coupled to the outer membrane via the ExbB-ExbD-TonB protein complex. Knowledge of the structure of this complex is required to elucidate the mechanisms of energy harvesting in the cytoplasmic membrane and energy transfer to the outer membrane transporters. Here we solubilized an ExbB oligomer and an ExbB-ExbD subcomplex from the cytoplasmic membrane with the detergent undecyl maltoside. Using laser-induced liquid bead ion desorption mass spectrometry (LILBID-MS), we determined at moderate desorption laser energies the oligomeric structure of ExbB to be mainly hexameric (ExbB(6)), with minor amounts of trimeric (ExbB(3)), dimeric (ExbB(2)), and monomeric (ExbB(1)) oligomers. Under the same conditions ExbB-ExbD formed a subcomplex consisting of ExbB(6)ExbD(1), with a minor amount of ExbB(5)ExbD(1). At higher desorption laser intensities, ExbB(1) and ExbD(1) and traces of ExbB(3)ExbD(1), ExbB(2)ExbD(1), ExbB(1)ExbD(1), ExbB(3), and ExbB(2) were observed. Since the ExbB(6) complex and the ExbB(6)ExbD(1) complex remained stable during solubilization and subsequent chromatographic purification on nickel-nitrilotriacetate agarose, Strep-Tactin, and Superdex 200, and during native blue gel electrophoresis, we concluded that ExbB(6) and ExbB(6)ExbD(1) are subcomplexes on which the final complex including TonB is assembled.  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.
(13)C NMR monitored the dynamics of exchange from specific hydrogens of hepatic [2-(13)C]glutamate and [3-(13)C]aspartate with deuterons from intracellular heavy water providing information on alpha-ketoglutarate/glutamate exchange and subcellular compartmentation. Mouse livers were perfused with [3-(13)C]alanine in buffer containing or not 50% (2)H(2)O for increasing periods of time (1 min < t < 30 min). Liver extracts prepared at the end of the perfusions were analyzed by high resolution (13)C NMR (150.13 MHz) with (1)H decoupling only and with simultaneous (1)H and (2)H decoupling. (13)C-(2)H couplings and (2)H-induced isotopic shifts observed in the glutamate C2 resonance, allowed to estimate the apparent rate constants (forward, reverse; min(-1)) for (i) the reversible exchange of [2-(13)C]glutamate H2 as catalyzed mainly by aspartate aminotransferase (0.32, 0.56), (ii) the reversible exchange of [2-(13)C]glutamate H3(proS) as catalyzed by NAD(P) isocitrate dehydrogenase (0.1, 0.05), and (iii) the irreversible exchanges of glutamate H3(proR) and H3(proS) as catalyzed by the sequential activities of mitochondrial aconitase and NAD isocitrate dehydrogenase of the tricarboxylic acid cycle (0.035), respectively. A similar approach allowed to determine the rates of (1)H-(2)H exchange for the H2 (0.4, 0.5) or H3(proR) (0.3, 0.2) or the H2 and H3(proS) hydrogens (0.20, 0.23) of [3-(13)C]aspartate isotopomers. The ubiquitous subcellular localization of (1)H-(2)H exchange enzymes and the exclusive mitochondrial localization of pyruvate carboxylase and the tricarboxylic acid cycle resulted in distinctive kinetics of deuteration in the H2 and either or both H3 hydrogens of [2-(13)C]glutamate and [3-(13)C]aspartate, allowing to follow glutamate and aspartate trafficking through cytosol and mitochondria.  相似文献   

16.
In order to determine the sequence dependence of the conformation of deoxynucleotides, Raman spectra have been obtained for the following oligodeoxynucleotides in aqueous salt solutions and in crystals: d(CpG)(I), d(TGCGCGCA)(II), d(CACGCGTG)(III), d(CGTGCACG)(IV), d(CGCATGCG)(V), d(ACGCGCGT)(VI), d(CGCGTACGCG)(VII), d(CGCACGTGCG)(VIII) and d(CGTGCGCACG)(IX), d(GCTATAGC) (X), d(GCATATGC) (XI), d(GGTATACC) (XII) and d(GGATATCC) (XIII). The normal B type conformation is observed for all the oligomer DNA's at low salt (0.1-1.0 M NaCl) concentration in the temperature range of 0-25 degrees C. It was considered possible that all of the first nine oligomers could go into the Z form in aqueous high salt (5.0-6.0 M NaCl) solutions, and under these conditions the last four were considered candidates to go into the A form. The B-type conformation was found to exist in high salt solutions for (I), (IV), (V), (VI), (X), (XI) and (XIII); the Z or partial Z conformation appears in high salt solution for the oligomers, (II), (III), (VII), (VIII) and (IX); an A or partial A conformation appears in high salt solution for (XII). In the crystalline state, (IV), (VIII), (X), and (XI) stay in the B-form and all of the other oligomers adopt the complete Z-form except for (XII) which crystallizes in the A form. In both the crystal and in aqueous solutions, the identification of the conformation genus was made by means of Raman spectroscopy. In the crystal of (I), grown at pH7.0, guanosine is found to be in C3'-endo/syn conformation and cytidine in C2'-endo/anti, which may be taken as the ideal building block of the typical Z conformation. At pH4, (I) crystallizes in a conformation similar to the B genus. A study of the thermally induced B to Z transition has been carried out for (II) and (III). Based on the analysis of Raman spectra of the alternating pyrimidine-purine oligomers which might be expected to go into the Z form, the tendency for these oligonucleotides to adopt the Z form can be ranked as: d(CGCGCGCG) greater than (II) greater than (III) greater than (V) approximately (VI) greater than (IV) for octamers and (VII) greater than (VIII) greater than (IX) for the decamers. Similarly, those oligomers which might have a tendency to go into the A form could be ranked as (XII) greater than (XIII) approximately (X) greater than (XI). These data should provide help in formulating rules for predicting the sequence dependence of the B to A and B to Z transitions. Some possible rules are explored, but precautions should be taken.  相似文献   

17.
Intra- and transcellular water movements in plants are regulated by the water permeability of the plasma membrane (PM) and vacuolar membrane (VM) in plant cells. In the present study, we investigated the osmotic water permeability of both PM (P ( f1)) and VM (P ( f2)), as well as the bulk osmotic water permeability of a protoplast (P ( f(bulk))) isolated from radish (Raphanus sativus) roots. The values of P ( f(bulk)) and P ( f2) were determined from the swelling/shrinking rate of protoplasts and isolated vacuoles under hypo- or hypertonic conditions. In order to minimize the effect of unstirred layer, we monitored dropping or rising protoplasts (vacuoles) in sorbitol solutions as they swelled or shrunk. P ( f1) was calculated from P ( f(bulk)) and P ( f2) by using the 'three-compartment model', which describes the theoretical relationship between P ( f1), P ( f2) and P ( f(bulk)) (Kuwagata and Murai-Hatano in J Plant Res, 2007). The time-dependent changes in the volume of protoplasts and isolated vacuoles fitted well to the theoretical curves, and solute permeation of PM and VM was able to be neglected for measuring the osmotic water permeability. High osmotic water permeability of more than 500 mum s(-1), indicating high activity of aquaporins (water channels), was observed in both PM and VM in radish root cells. This method has the advantage that P ( f1) and P ( f2) can be measured accurately in individual higher plant cells.  相似文献   

18.
The reaction mechanism for selective oxidation of propylamine on oxygen-covered gold has been studied by the density functional theory (DFT) and generalized gradient approximation (GGA) with slab model. Our calculation results indicated that the adsorption energy of propylamine decreases with the increasing oxygen coverage, that is -0.38, -0.20 and -0.10 eV on clean, 2/9 monolayer (ML) and 2/3 monolayer (ML) oxygen, respectively. The adsorption energies of the intermediates also have the trend of the gradual lower. The present work also indicated that the final product distribution depends on the oxygen coverage: propylamine undergoes N-H bond and C-H bond cleavage to produce propionitrile and water at low-oxygen-coverage (θ(o)?=?2/9 ML), and to yield propionitrile, propionaldehyde and water at high-oxygen-coverage (θ(o)?=?2/3 ML). The energy barrier of the first step of propyamine oxidation (CH(3)CH(2)CH(2)NH(2)?→?CH(3)CH(2)CH(2)NH) is 0.16 eV (θ(o)?=?2/9 ML) and 0.38 eV (θ(o)?=?2/3 ML). On the second step, the barrier energy is 0.16 (θ(o)?=?2/9 ML) and 0.25 (θ(o)?=?2/3 ML) eV of CH(3)CH(2)CH(2)NH?→?CH(3)CH(2)CH(2)N, next both C-H breakage and the barrier energy is 0.20 eV (CH(3)CH(2)CH(2)N?→?CH(3)CH(2)CHN) and 0.25 eV (CH(3)CH(2)CHN?→?CH(3)CH(2)CN) on low oxygen coverage, and 0.15 eV (CH(3)CH(2)CH(2)N?→?CH(3)CH(2)CHN) and 0.26 eV(CH(3)CH(2)CHN?→?CH(3)CH(2)CN) on the high oxygen coverage. The additional reaction step of CH(3)CH(2)CHN?→?CH(3)CH(2)CHO occurs on the high oxygen coverage, and the associated barrier is 0.41 eV. The calculation results show that the oxidation of propylamine can occur at room temperature due to the lower energy barrier. Furthermore, it was found that the energy barrier for the possible reaction steps at the low oxygen coverage is generally smaller than that on high oxygen coverage, which agrees with the experimental results.  相似文献   

19.
The binding of carbon dioxide by horse haemoglobin   总被引:15,自引:7,他引:8  
1. Three modified horse haemoglobins have been prepared: (i) alpha(c) (2)beta(c) (2), in which both the alpha-amino groups of the alpha- and beta-chains have reacted with cyanate, (ii) alpha(c) (2)beta(2), in which the alpha-amino groups of the alpha-chains have reacted with cyanate, and (iii) alpha(2)beta(c) (2), in which the two alpha-amino groups of the beta-chain have reacted with cyanate. 2. The values of n (the Hill constant) for alpha(c) (2)beta(c) (2), alpha(2)beta(c) (2) and alpha(c) (2)beta(2) were (respectively) 2.5, 2.0 and 2.6, indicating the presence of co-operative interactions between the haem groups for all derivatives. 3. In the alkaline pH range (about pH8.0) all the derivatives show the same charge as normal haemoglobin whereas in the acid pH range (about pH6.0) alpha(c) (2)beta(c) (2) differs by four protonic charges and alpha(c) (2)beta(2), alpha(2)beta(c) (2) by two protonic charges from normal haemoglobin, indicating that the expected number of ionizing groups have been removed. 4. alpha(c) (2)beta(2) and alpha(c) (2)beta(c) (2) show a 25% decrease in the alkaline Bohr effect, in contrast with alpha(2)beta(c) (2), which has the same Bohr effect as normal haemoglobin. 5. The deoxy form of alpha(c) (2)beta(c) (2) does not bind more CO(2) than the oxy form of alpha(c) (2)beta(c) (2), whereas alpha(c) (2)beta(2) and alpha(2)beta(c) (2) show intermediate binding. 6. The results reported confirm the hypothesis that, under physiological conditions, haemoglobin binds CO(2) through the four terminal alpha-amino groups and that the two terminal alpha-amino groups of alpha-chains are involved in the Bohr effect.  相似文献   

20.
We studied the seasonal fluctuation of soil respiration (R(S)), and its root-dependent (R(R)) and basal (R(B)) components, in a Vitis vinifera (Chardonnay) vineyard. The R(S) components were estimated through independent field methods (y-intercept and trenching) and modeled on the basis of a Q(10) response to soil temperature, and fine and coarse root respiration coefficients. The effect of assimilate availability on R(R) was assessed through a trunk girdling treatment. The apparent Q(10) for R(R) was twice that of R(B) (3.5 vs 1.6) and increased linearly with increasing vine root biomass. The fastest R(R) of fine roots was during rapid fruit growth and the fastest R(R) of coarse roots was immediately following fruit development. R(S) was estimated at 32.6 kg ha(-1) d(-1) (69% as a result of R(R) ) for the hottest month and at 7.6 kg ha(-1) d(-1) (18% as a result of R(R)) during winter dormancy. Annual R(S) was low compared with other natural and cultivated ecosystems: 5.4 Mg ha(-1) (46% as a result of R(R)). Our estimates of annual vineyard R(S) are the first for any horticultural crop and suggest that the assumption that they are similar to those of annual crops or forest trees might lead to an overestimation.  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