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.
Nitrophorins 1-4 (NP1-4) are ferriheme proteins from the blood-sucking insect Rhodnius prolixus that transport nitric oxide (NO) to the victim, sequester histamine, and inhibit blood coagulation. Here, we report kinetic and thermodynamic analyses for ligand binding by all four proteins and their reduction potentials. All four undergo biphasic association and dissociation reactions with NO. The initial association is fast (1.5-33 microM(-)(1) s(-)(1)) and similar to that of elephant metmyoglobin. However, unlike in metmyoglobin, a slower second phase follows ( approximately 50 s(-)(1)), and the stabilized final complexes are resistant to autoreduction (E degrees = +3 to +154 mV vs normal hydrogen electrode). NO dissociation begins with a slow, pH-dependent step (0.02-1.4 s(-)(1)), followed by a faster phase that is again similar to that of metmyoglobin (3-52 s(-)(1)). The equilibrium dissociation constants are quite small (1-850 nM). NP1 and NP4 display larger release rate constants and smaller association rate constants than NP2 and NP3, leading to values for K(d) that are about 10-fold greater. The results are discussed in light of the recent crystal structures of NP1, NP2, and NP4, which display open, polar distal pockets, and of NP4-NO, which displays an NO-induced conformational change that leads to expulsion of solvent and complete burial of the NO ligand in a now nonpolar distal pocket. Taken together, the results suggest that tighter NO binding in the nitrophorins is due to the trapping of the molecule in a nonpolar distal pocket rather than through formation of particularly strong Fe-NO or hydrogen bonds.  相似文献   

2.
The [Ru(II)(Hedta)NO(+)] complex is a diamagnetic species crystallizing in a distorted octahedral geometry, with the Ru-N(O) length 1.756(4) A and the RuNO angle 172.3(4) degrees . The complex contains one protonated carboxylate (pK(a)=2.7+/-0.1). The [Ru(II)(Hedta)NO(+)] complex undergoes a nitrosyl-centered one-electron reduction (chemical or electrochemical), with E(NO+/NO)=-0.31 V vs SCE (I=0.2 M, pH 1), yielding [Ru(II)(Hedta)NO](-), which aquates slowly: k(-NO)=2.1+/-0.4x10(-3) s(-1) (pH 1.0, I=0.2 M, CF(3)COOH/NaCF(3)COO, 25 degrees C). At pHs>12, the predominant species, [Ru(II)(edta)NO](-), reacts according to [Ru(II)(edta)NO](-)+2OH(-)-->[Ru(II)(edta)NO(2)](3-), with K(eq)=1.0+/-0.4 x 10(3) M(-2) (I=1.0 M, NaCl; T=25.0+/-0.1 degrees C). The rate-law is first order in each of the reactants for most reaction conditions, with k(OH(-))=4.35+/-0.02 M(-1)s(-1) (25.0 degrees C), assignable mechanistically to the elementary step comprising the attack of one OH(-) on [Ru(II)(edta)NO](-), with subsequent fast deprotonation of the [Ru(II)(edta)NO(2)H](2-) intermediate. The activation parameters were DeltaH(#)=60+/-1 kJ/mol, DeltaS(#)=-31+/-3 J/Kmol, consistent with a nucleophilic addition process between likely charged ions. In the toxicity up-and-down tests performed with Swiss mice, no death was observed in all the doses administered (3-9.08 x 10(-5) mol/kg). The biodistribution tests performed with Wistar male rats showed metal in the liver, kidney, urine and plasma. Eight hours after the injection no metal was detected in the samples. The vasodilator effect of [Ru(II)(edta)NO](-) was studied in aortic rings without endothelium, and was compared with sodium nitroprusside (SNP). The times of maximal effects of [Ru(II)(edta)NO](-) and SNP were 2 h and 12 min, respectively, suggesting that [Ru(II)(edta)NO](-) releases NO slowly to the medium in comparison with SNP.  相似文献   

3.
A number of ferriheme proteins, termed nitrophorins (NPs), occur in the saliva of the bloodsucking insect Rhodnius prolixus ('kissing bug'), which is a vector for Chagas' disease. Nitrophorins bind the heme b cofactor in the beta-barrel of their lipocalin fold, which is further anchored through a proximal histidine-Fe(III) bond. The distal Fe(III) coordination site then binds nitric oxide (NO) for delivery into a host's tissues during blood feeding, where, upon NO release, the distal Fe(III) site acts as a histamine trap to delay the victim's immune response. Previously, four nitrophorins from R. prolixus, NP1 to NP4, have been extensively characterized. Recently, another nitrophorin, NP7, was discovered in a cDNA library derived from the same insect. Among the R. prolixus nitrophorins, NP7 was found to be unique in its ability to bind to negatively charged cell surfaces. However, the yield of functional recombinant NP7 was rather low when the established protocol for NP1-4 was followed. Here, we report on a novel expression and reconstitution method for NP7 that yields sufficient amounts of pure protein for extensive characterization (28-fold increase). This method may prove useful for the reconstitution of other proteins with a lipocalin fold.  相似文献   

4.
The saliva of blood-feeding insects contains a variety of molecules having antihemostatic activity. Here, we describe nitrophorin 7 (NP7), a salivary protein that binds with high affinity to anionic phospholipid membranes. The protein is apparently targeted to the negatively charged surfaces of activated platelets and other cells, where it can serve as a vasodilator, antihistamine, platelet aggregation inhibitor, and anticoagulant. As with other members of the nitrophorin group, NP7 reversibly binds a molecule of NO and binds histamine with high affinity. The protein differs from other nitrophorins in that it binds to membranes containing phosphatidylserine. Sedimentation and surface plasmon resonance experiments, revealed two classes of phospholipid-binding sites having K(d) values of 4.8 and 755 nM. NP7 inhibits prothrombin activation by blocking phospholipid binding sites for the prothrombinase complex on the surfaces of vesicles and activated platelets. As a NO complex, NP7 inhibits collagen and ADP-induced platelet aggregation and induces disaggregation of ADP-stimulated platelets by an NO-mediated mechanism. Molecular modeling of NP7 revealed a putative, positively charged membrane interaction surface comprised mainly of a helix lying outside of the lipocalin beta-barrel structure.  相似文献   

5.
By application of pulse radiolysis it was demonstrated that nitrogen dioxide (NO2.) oxidizes Gly-Tyr in aqueous solution with a strongly pH-dependent rate constant (k6 = 3.2 X 10(5) M-1 S-1 at pH 7.5 and k6 = 2.0 X 10(7) M-1 S-1 at pH 11.3), primarily generating phenoxyl radicals. The phenoxyl can react further with NO2. (k7 approximately 3 X 10(9) M-1 S-1) to form nitrotyrosine, which is the predominant final product in neutral solution and at low tyrosyl concentrations under gamma-radiolysis conditions. Tyrosine nitration is less efficient in acidic solution, due to the natural disproportionation of NO2., and in alkaline solutions and at high tyrosyl concentrations due to enhanced tyrosyl dimerization. Selective tyrosine nitration by interaction of NO2. with proteins (at pH 7 to 9) was demonstrated in the case of histone, lysozyme, ribonuclease A, and subtilisin Carlsberg. Nitrotyrosine developed slowly also under incubation of Gly-Tyr with nitrite at pH 4 to 5, where NO2. is formed by acid decomposition of HONO. It is recalled in this context that NO2.-induced oxidations, by regenerating NO2-, can propagate NO2./NO2- redox cycling under acidic conditions. Even faster than with tyrosine is the NO2.-induced oxidation of cysteine-thiolate (k9 = 2.4 X 10(8) M-1 S-1 at pH 9.2), involving the transient formation of cystinyl radical anions. The interaction of NO2. with Gly-Trp was comparably slow (k approximately 10(6) M-1 S-1), and no reaction was detectable by pulse radiolysis with Met-Gly and (Cys-Gly)2, or with DNA. Slow reactions of NO2. were observed with arachidonic acid (k approximately 10(6) M-1 S-1 at pH 9.0) and with linoleate (k approximately 2 X 10(5) M-1 S-1 at pH 9.4), indicating that NO2. is capable of initiating lipid peroxidation even in an aqueous environment. NO2.-Induced tyrosine nitration, using 50 microM Gly-Tyr at pH 8.2, was hardly inhibited, however, in the presence of 1 mM linoleate, and was not affected at all in the presence of 5 mM dimethylamine (a nitrosamine precursor). It is concluded that protein modifications, and particularly phenol and thiol oxidation, may be an important mechanism, as well as initiation of lipid peroxidation, of action of NO2. in biological systems.  相似文献   

6.
Rhodnius prolixus is a blood-sucking insect that uses a mixture of nitrophorin (NP) proteins to deliver nitric oxide (NO) from the insect saliva to the hosts via a ferric heme coordinated to the protein, causing vasodilatation and anticoagulation to support their feeding. R. prolixus NPs 1-4 are very similar proteins ( approximately 20 kDa) with different NO affinities for stepwise NO release triggered by pH increase and histamine binding in hosts. Ultra-high-resolution X-ray structures of native and mutant NPs and their kinetic analysis already have revealed the fundamental steps of NO binding and release. In this study, we found that NPs can exist in multiple oligomerization states at higher concentrations. The oligomers are characterized by a combination of multiple biophysical methods. The intrinsic features of the oligomerization revealed here led us to propose that this intensive, moderately pH- and ligand-dependent oligomerization of NPs has physiological implications in the facilitation of the efficient storage and release of the highly reactive NO in the insect saliva and the victim, respectively.  相似文献   

7.
One physiological significance of the red blood cell (RBC) structure is that NO binding of Hb is retarded by encapsulation with the cell membrane. To clarify the mechanism, we analyzed Hb-vesicles (HbVs) with different intracellular Hb concentrations, [Hb](in), and different particle sizes using stopped-flow spectrophotometry. The apparent NO binding rate constant, k(on)('(NO)), of HbV at [Hb](in) = 1 g/dl was 2.6 x 10(7) m(-1) s(-1), which was almost equal to k(on)((NO)) of molecular Hb, indicating that the lipid membrane presents no obstacle for NO binding. With increasing [Hb](in) to 35 g/dl, k(on)('(NO)) decreased to 0.9 x 10(7) m(-1) s(-1), which was further decreased to 0.5 x 10(7) m(-1) s(-1) with enlarging particle diameter from 265 to 452 nm. For CO binding, which is intrinsically much slower than NO binding, k(on)('(CO)) did not change greatly with [Hb](in) and the particle diameter. Results obtained using diffusion simulations coupled with elementary binding reactions concur with these tendencies and clarify that NO is trapped rapidly by Hb from the interior surface region to the core of HbV at a high [Hb](in), retarding NO diffusion toward the core of HbV. In contrast, slow CO binding allows time for further CO-diffusion to the core. Simulations extrapolated to larger particles (8 mum) showing retardation even for CO binding. The obtained k(on)('(NO)) and k(on)('(NO)) yield values similar to those reported for RBCs. In summary, the intracellular, not extracellular, diffusion barrier is predominant due to the rapid NO binding that induces a rapid sink of NO from the interior surface to the core, retarding further NO diffusion and binding.  相似文献   

8.
The amino acid sequence of human myoglobin (Mb) is similar to other mammalian Mb except for a unique cysteine residue at position 110 (Cys(110)). Anaerobic treatment of ferrous forms of wild-type human Mb, the C110A variant of human Mb or horse heart Mb, with either authentic NO or chemically derived NO in vitro yields heme-NO complexes as detected by electron paramagnetic resonance spectroscopy (EPR). By contrast, no EPR-detectable heme-NO complex was observed from the aerobic reactions of NO and either the ferric or oxy-Mb forms of wild-type human or horse heart myoglobins. Mass analyses of wild-type human Mb treated aerobically with NO indicated a mass increase of approximately 30 atomic mass units (i.e., NO/Mb = 1 mol/mol). Mass analyses of the corresponding apoprotein after heme removal showed that NO was associated with the apoprotein fraction. New electronic maxima were detected at A(333 nm) (epsilon = 3665 +/- 90 mol(-)(1) cm(-)(1); mean +/- S.D.) and A(545 nm) (epsilon = 44 +/- 3 mol(-)(1) cm(-)(1)) in solutions of S-nitrosated wild-type human Mb (similar to S-nitrosoglutathione). Importantly, the sulfhydryl S-H stretch vibration for Cys(110) measured by Fourier transform infrared (nu approximately 2552 cm(-)(1)) was absent for both holo- and apo- forms of the wild-type human protein after aerobic treatment of the protein with NO. Together, these data indicate that the reaction of wild-type human Mb and NO yields either heme-NO or a novel S-nitrosated protein dependent on the oxidation state of the heme iron and the presence or absence of dioxygen.  相似文献   

9.
The objective of this study was to elucidate the origin of the nitric oxide-forming reactions from nitrite in the presence of the iron-N-methyl-D-glucamine dithiocarbamate complex ((MGD)(2)Fe(2+)). The (MGD)(2)Fe(2+) complex is commonly used in electron paramagnetic resonance (EPR) spectroscopic detection of NO both in vivo and in vitro. Although it is widely believed that only NO can react with (MGD)(2)Fe(2+) complex to form the (MGD)(2)Fe(2+).NO complex, a recent article reported that the (MGD)(2)Fe(2+) complex can react not only with NO, but also with nitrite to produce the characteristic triplet EPR signal of (MGD)(2)Fe(2+).NO (Hiramoto, K., Tomiyama, S., and Kikugawa, K. (1997) Free Radical Res. 27, 505-509). However, no detailed reaction mechanisms were given. Alternatively, nitrite is considered to be a spontaneous NO donor, especially at acidic pH values (Samouilov, A., Kuppusamy, P., and Zweier, J. L. (1998) Arch Biochem. Biophys. 357, 1-7). However, its production of nitric oxide at physiological pH is unclear. In this report, we demonstrate that the (MGD)(2)Fe(2+) complex and nitrite reacted to form NO as follows: 1) (MGD)(2)Fe(2).NO complex was produced at pH 7.4; 2) concomitantly, the (MGD)(3)Fe(3+) complex, which is the oxidized form of (MGD)(2)Fe(2+), was formed; 3) the rate of formation of the (MGD)(2)Fe(2+).NO complex was a function of the concentration of [Fe(2+)](2), [MGD], [H(+)] and [nitrite].  相似文献   

10.
The kinetics of the reversible binding of cyanide by the ferric cytochrome c' from Chromatium vinosum have been studied over the pH range 6.9-9.6. The reaction is extremely slow at neutral pH compared to the reactions of other high-spin ferric heme proteins with cyanide. The observed bimolecular rate constant at pH 7.0 is 2.25 X 10(-3) M-1 s-1, which is approximately 10(7)-fold slower than that for peroxidases, approximately 10(5)-fold slower than those for hemoglobin and myoglobin, and approximately 10(2)-fold to approximately 10(3)-fold slower than that recently reported for the Glycera dibranchiata hemoglobin, which has anomalously slow cyanide rate constants of 4.91 X 10(-1), 3.02 X 10(-1), and 1.82 M-1 s-1 for components II, III, and IV, respectively [Mintorovitch, J., & Satterlee, J. D. (1988) Biochemistry 27, 8045-8050; Mintorovitch, J., Van Pelt, D., & Satterlee, J. D. (1989) Biochemistry 28, 6099-6104]. The unusual ligand binding property of this cytochrome c' is proposed to be associated with a severely hindered heme coordination site. Cyanide binding is also characterized by a nonlinear cyanide concentration dependence of the observed rate constant at higher pH values, which is interpreted as involving a change in the rate-determining step associated with the formation of an intermediate complex between the cytochrome c' and cyanide prior to coordination. The pH dependence of both the binding constant for the formation of the intermediate complex and the association rate constant for the subsequent coordination to the heme can be attributed to the ionization of HCN, where cyanide ion binding is the predominant process.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The surface loop which in the Bacillus subtilis neutral protease (NP) extends from amino acid residue 188 to residue 194 was replaced, by site-directed mutagenesis, with the 10-residue segment which in the homologous polypeptide chain of thermolysin (TLN) binds calcium-4 [Matthews, B. W., Weaver, L. H., & Kester, W. R. (1974) J. Biol. Chem. 249, 8030-8044]. The mutant NP was isolated to homogeneity, and its structural, functional, calcium-binding, and stability properties were investigated. Proteolytic fragmentation with Staphylococcus aureus V8 protease of mutant NP was used to isolate and analyze the protein fragment encompassing the site of mutation, unambiguously establishing the effective insertion of the new 10-residue segment. Atomic absorption measurements allowed us to demonstrate that mutant NP binds three calcium ions instead of the two ions bound to wild-type NP, showing that indeed the chain segment grafted from TLN to NP maintains its calcium-binding properties. The mutant NP showed kinetic parameters essentially similar to those of the wild-type NP with Z-Phe-Leu-Ala-OH as substrate. The enzyme inactivation of mutant vs wild-type NP was studied as a function of free [Ca2+]. It was found that mutant NP was much less stable than the wild-type NP when enzyme solutions were dialyzed at neutral pH in the presence of [Ca2+] below 10(-3) M. On the other hand, the kinetic thermal stability to irreversible inactivation of mutant NP, when measured in the presence of 0.1 M CaCl2, was found to be increased about 2-fold over that of the wild-type NP. Thus, modulation of enzyme stability by free [Ca2+] in mutant NP correlates with similar findings previously reported for thermolysin. Overall, the results obtained indicate that protein engineering experiments can be used to prepare hybrid proteins on the basis of sequence and function analysis of homologous protein molecules and show the feasibility of engineering metal ion binding sites into proteins.  相似文献   

12.
Recent mathematical models of bivalent hapten-induced histamine release from basophils predict that under appropriate conditions histamine release is maximum when cross-link formation is maximum, at a hapten concentration equal to 1/(2Ka), where Ka is the average affinity constant of the hapten for a single IgE binding site. To test this prediction we sensitized human basophils with a monoclonal anti-dinitrophenol IgE and generated histamine release dose-response curves with a bivalent hapten, alpha, epsilon-DNP-lysine. The monoclonal IgE has a published affinity constant of 7.1 X 10(7) M-1 for epsilon-DNP-lysine as determined by equilibrium dialysis. From the position of the maximum of the histamine dose-response curves, both in the presence and in the absence of monovalent DNP hapten, we determine that the sensitizing IgE has an intrinsic affinity constant of 6.9 +/- 0.5 X 10(7) M-1 for epsilon-DNP-lysine and 1.2 +/- 0.6 X 10(6) M-1 for alpha-DNP-lysine. The agreement between the two estimates of the epsilon-DNP-lysine affinity constant, one from histamine release experiments involving surface bound IgE and one from binding experiments involving IgE free in solution, 1) is consistent with a central prediction of the theory of cross-linking and 2) indicates that the hapten-binding properties of the IgE are unaffected by its being bound to Fc epsilon receptors on the basophil surface.  相似文献   

13.
We have designed and synthesized new optically active bisviologens ([BNMV](4+)) containing a binaphthyl moiety to examine the stereoselective photoinduced electron-transfer (ET) reactions with zinc-substituted myoglobin (ZnMb) by flash photolysis. The photoexcited triplet state of ZnMb, (3)(ZnMb)*, was successfully quenched by [BNMV](4+) ions to form the radical pair of a ZnMb cation (ZnMb(.+)) and a reduced viologen ([BNMV](.3+)), followed by a thermal ET reaction to the ground state. The rate constants ( k(q)) for the ET quenching at 25 degrees C were obtained as k(q)( R)=(2.9+/-0.2)x10(7) M(-1) s(-1) and k(q)( S)=(2.2+/-0.2)x10(7) M(-1) s(-1), respectively. The ratio of k(q)( R)/ k(q)( S)=1.3 indicates that the ( R)-isomer of the chiral viologen preferentially quenches (3)(ZnMb)*. On the other hand, the rate constants ( k) for the thermal ET reaction from [BNMV](.3+) to ZnMb(-+) at 25 degrees C were k( R)=(1.2+/-0.1)x10(8) M(-1) s(-1) and k( S)=(0.47+/-0.03)x10(8) M(-1) s(-1), respectively, and the ratio remarkably increased to k( R)/ k( S)=2.6. The activation parameters, Delta H(not equal) and Delta S(not equal), were determined from the kinetic measurements at various temperatures (10-30 degrees C) to understand the ET mechanisms. In the quenching reaction, the energy differences of Delta Delta H*(R- S) and T Delta Delta S*( R- S) at 25 degrees C were calculated to be -3.9+/-1.6 and -3.3+/-0.2 kJ mol(-1), respectively, whereas Delta Delta H*( R-S)=7.7+/-1.9 kJ mol(-1 )and T Delta Delta S*( R-S)=9.9+/-0.5 kJ mol(-1 )were found for the thermal ET reaction. Therefore, the thermal ET reaction to the ground state was proved to be dominated by the entropy term, and the large stereoselectivity may arise from the decrease in charge repulsion between donor and acceptor.  相似文献   

14.
Cytochrome P-450(M-1) [P-450(M-1)] is specifically expressed in adult male rat liver [Yoshioka, H., Morohashi, K., Sogawa, K., Miyata, T., Kawajiri, K., Hirose, T., Inayama, S., Fujii-Kuriyama, Y., & Omura, T. (1987) J. Biol. Chem. 262, 1706-1711]. Isolation and analysis of the gene for P-450(M-1) revealed that the coding region of the gene is interrupted by eight introns and is dispersed over a 35-kilobase pair region of chromosomal DNA. Intron insertion sites of the P-450(M-1) gene are located at equivalent positions to those of cytochrome P-450b and P-450e, which are phenobarbital-inducible. Sequence analysis of the 5'-upstream region of the P-450(M-1) gene shows that there is a homologous sequence to glucocorticoid regulatory elements (GRE) identified in glucocorticoid-responsive genes.  相似文献   

15.
The dissociation of nitric oxide from hemoglobin, from isolated subunits of hemoglobin, and from myoglobin has been studied using dithionite to remove free nitric oxide. The reduction of nitric oxide by dithionite has a rate of 1.4 X 10(3) M-1 S-1 at 20 degrees in 0.05 M phosphate, pH 7.0, which is small compared with the rate of recombination of hemoglobin with nitric oxide (25 X 10(6) M-1 S-1 (Cassoly, R., and Gibson, Q. H. (1975) J. Mol. Biol. 91, 301-313). The rate of NO combination with chains and myoglobin was found to be 24 X 10(6) M-1 S-1 and 17 X 10(6) M-1 S-1, respectively. Hence, the observed progress curve of the dissociation of nitric oxide is dependent upon the dithionite concentration and the total heme concentration. Addition of excess carbon monoxide to the dissociation mixture reduces the free heme yielding a single exponential process for chains and for myoglobin which is dithionite and heme concentration independent over a wide range of concentrations. The rates of dissociation of nitric oxide from alpha chains, from beta chains, and from myoglobin are 4.6 X 10(-5) S-1, 2.2 X 10(-5) S-1, and 1.2 X 10(4) S-1, respectively, both in the presence and in the absence of carbon monoxide at 20 degrees in 0.05 M phosphate, pH 7.0. Analogous heme and dithionite concentration dependence is found for the dissociation of nitric oxide from tetrameric hemoglobin. The reaction is cooperative, the intrinsic rate constants for the dissociation of the 1st and 4th molecules of NO differing about 100-fold. With hemoglobin, replacement of NO by CO at neutral pH is biphasic in phosphate buffers. The rate of the slow phase is 1 X 10(-5) S-1 and is independent of pH. The amplitude of the fast phase increases with lowering of pH. By analogy with the treatment of the HbCO + NO reaction given by Salhany et al. (Salhany, J.M., Ogawa, S., and Shulman, R.G. (1975) Biochemistry 14, 2180-2190), the fast phase is attributed to the dissociation of NO from T state molecules and the slow phase to dissociation from R state molecules. Analysis of the data gives a pH-independent value of 0.01 for the allosteric constant c (c = Kr/Kt where Kr and Kt are the dissociation constants for NO from the R and T states, respectively) and pH-dependent values of L (2.5 X 10(7) at pH 7 in 0.05 M phosphate buffer). The value of c is considerably greater than that for O2 and CO. Studies of the difference spectrum induced in the Soret region by inositol hexaphosphate are also reported. This spectrum does not arise directly from the change of conformation between R and T states. The results show that if the equilibrium binding curve for NO could be determined experimentally, it would show cooperativity with Hill's n at 50% saturation of about 1.6.  相似文献   

16.
The binding of bilirubin and the polyene fatty acids cis-parinaric acid and cis-eleostearic acid to human alpha-fetoprotein was studied using fluorescence quenching and fluorescence enhancement techniques. alpha-Fetoprotein has three fatty acid binding sites of decreasing affinity (association constants 2.1 x 10(7) M-1 9.1 X 10(5) M-1, and 1.4 x 10(5) M-1) and one relatively strong and one relatively weak bilirubin binding site (association constants 1.1 x 10(7) M-1 and 1.8 x 10(5) M-1). These association constants are slightly weaker than the corresponding association constants for binding to human albumin. Competition experiments failed to show preferential binding of polyunsaturated fatty acids. Fluorescence quenching was used to determine 11 ligand-ligand and ligand-tryptophanyl residue distances. Each of these 11 calculated distances (ranging from 19 A to 32 A) was within 5 A of the corresponding distances measured previously for human albumin (Berde, C.B., Hudson, B.S., Simoni, R.D., and Sklar, L.A. 1979, J. Biol. Chem. 254, 391-400). Thus, in addition to previously described sequence homology, immunologic cross-reactivity, and other similarities, human albumin and human alpha-fetoprotein have functional and geometric homologies.  相似文献   

17.
Spectroscopic and electrochemical study of the [Fe(4)(mu(3)-S)(3)(NO)(7)](-) photochemical reaction and thermodynamic calculations of relevant systems demonstrate the redox character of this process. The photoinduced electron transfer between substrate clusters in excited and ground state (probably via exciplex formation) results in dismutation yielding unstable [Fe(4)(mu(3)-S)(3)(NO)(7)](2-) and [Fe(4)(mu(3)-S)(3)(NO)(7)](0). Back electron transfer between the primary products is responsible for fast reversibility of the photochemical reaction in deoxygenated solutions. In the presence of an electron acceptor (such as O(2), MV(2+) or NO) an oxidative quenching of the (*)[Fe(4)(mu(3)-S)(3)(NO)(7)](-) is anticipated, although NO seems to participate as well in the reductive quenching. The electron acceptors can also regenerate the substrate from its reduced form ([Fe(4)(mu(3)-S)(3)(NO)(7)](2-)), whereas the other primary product ([Fe(4)(mu(3)-S)(3)(NO)(7)](0)) decomposes to the final products. The suggested mechanism fits well to all experimental observations and shows the thermodynamically favored pathways and explains formation of all major (Fe(2+), S(2-), NO) and minor products (N(2)O, Fe(3+)). The photodissociation of nitrosyl ligands suggested earlier as the primary photochemical step cannot be, however, definitely excluded and may constitute a parallel pathway of [Fe(4)(mu(3)-S)(3)(NO)(7)](-) photolysis.  相似文献   

18.
Histamine stimulated the accumulation of [3H]inositol 1-phosphate in the presence of 10 mM LiCl in [3H]inositol-loaded tissue slices from several regions of guinea pig brain. The level of [3H]inositol 1-phosphate increased approximately linearly, after an initial lag period, up to a time of 120 min. In the absence of lithium ions the accumulation of the 1-phosphate stimulated by histamine in cerebral cortical and hippocampal slices was markedly reduced. Lithium ions had much less effect on the response to histamine in cerebellar slices. The characteristics of the response to histamine were consistent with mediation by H1 receptors, and the affinity constants derived for mepyramine (2.3 X 10(9) M-1) and methapyrilene (1.8 X 10(8) M-1) were similar to those reported from measurements on other H1 responses in the guinea pig. The EC50 for histamine was similar in cerebellum, cerebral cortex, hippocampus, and hypothalamus. The position of the dose-response curve for histamine in cerebral cortical slices was similar to that of the curve for the receptor binding of histamine deduced from histamine inhibition of [3H]mepyramine binding.  相似文献   

19.
The interaction of nitric oxide (NO) with iron-sulfur cluster proteins results in the formation of dinitrosyl iron complexes (DNICs) coordinated by cysteine residues from the peptide backbone or with low molecular weight sulfur-containing molecules like glutathione. Such DNICs are among the modes available in biology to store, transport, and deliver NO to its relevant targets. In order to elucidate the fundamental chemistry underlying the formation of DNICs and to characterize possible intermediates in the process, we have investigated the interaction of NO (g) and NO(+) with iron-sulfur complexes having the formula [Fe(SR)(4)](2-), where R=(t)Bu, Ph, or benzyl, chosen to mimic sulfur-rich iron sites in biology. The reaction of NO (g) with [Fe(S(t)Bu)(4)](2-) or [Fe(SBz)(4)](2-) cleanly affords the mononitrosyl complexes (MNICs), [Fe(S(t)Bu)(3)(NO)](-) (1) and [Fe(SBz)(3)(NO)](-) (3), respectively, by ligand displacement. Mononitrosyl species of this kind were previously unknown. These complexes further react with NO (g) to generate the corresponding DNICs, [Fe(SPh)(2)(NO)(2)](-) (4) and [Fe(SBz)(2)(NO)(2)](-) (5), with concomitant reductive elimination of the coordinated thiolate donors. Reaction of [Fe(SR)(4)](2-) complexes with NO(+) proceeds by a different pathway to yield the corresponding dinitrosyl S-bridged Roussin red ester complexes, [Fe(2)(mu-S(t)Bu)(2)(NO)(4)] (2), [Fe(2)(mu-SPh)(2)(NO)(4)] (7) and [Fe(2)(mu-SBz)(2)(NO)(4)] (8). The NO/NO(+) reactivity of an Fe(II) complex with a mixed nitrogen/sulfur coordination sphere was also investigated. The DNIC and red ester species, [Fe(S-o-NH(2)C(6)H(4))(2)(NO)(2)](-) (6) and [Fe(2)(mu-S-o-NH(2)C(6)H(4))(2)(NO)(4)] (9), were generated. The structures of 8 and 9 were verified by X-ray crystallography. The MNIC complex 1 can efficiently deliver NO to iron-porphyrin complexes like [Fe(TPP)Cl], a reaction that is aided by light. Removal of the coordinated NO ligand of 1 by photolysis and addition of elemental sulfur generates higher nuclearity Fe/S clusters.  相似文献   

20.
G Simon  J Filep  T Zelles 《Life sciences》1990,47(22):2021-2025
Alpha adrenergic agonists and antagonists as clonidine, guanfacine, yohimbine, phenylephrine and prazosin inhibited the [3H]-QNB binding to rat brain cortex muscarinic acetylcholine receptor (mAChR, M-1 subtype), heart (M-2 subtype) and parotid gland homogenate (M-3 subtype) in a dose-dependent competitive fashion. Ki values were between 10(-6) and 10(-3) M. Hill coefficients were about 1. No correlation was found between mAChR inhibiting capacity of these drugs and their activity on alpha adrenergic receptors. In contrast, other transmitters, as dopamine, GABA, glutamic acid, histamine, serotonin, isoproterenol and platelet activating factor (PAF) did not affect the QNB binding.  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