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 15 毫秒
1.
The interaction of the adenylate cyclase catalytic domain (AC) of the Bordetella pertussis major exotoxin with its activator calmodulin (CaM) was studied by time-resolved fluorescence spectroscopy using three fluorescent groups located in different regions of AC: tryptophan residues (W69 and W242), a nucleotide analogue (3'-anthraniloyl-2'-deoxyadenosine 5'-triphosphate, Ant-dATP) and a cysteine-specific probe (acrylodan). CaM binding elicited large changes in the dynamics of W242, which dominates the fluorescence emission of both AC and AC-CaM, similar to that observed for isolated CaM-binding sequences of different lengths [Bouhss, A., Vincent, M., Munier, H., Gilles, A.M., Takahashi, M., Barzu, O., Danchin, A. & Gallay, J. (1996) Eur. J. Biochem.237, 619-628]. In contrast, Ant-dATP remains completely immobile and inaccessible to the solvent in both the AC and AC-CaM nucleotide-binding sites. As AC contains no cysteine residue, a single-Cys mutant at position 75 was constructed which allowed labeling of the catalytic domain with acrylodan. Its environment is strongly apolar and rigid, and only slightly affected by CaM. The protein's hydrodynamic properties were also studied by fluorescence anisotropy decay measurements. The average Brownian rotational correlation times of AC differed significantly according to the probe used (19 ns for W242, 25 ns for Ant-dATP, and 35 ns for acrylodan), suggesting an elongated protein shape (axial ratio of approximately 1.9). These values increased greatly with the addition of CaM (39 ns for W242, 60-70 ns for Ant-dATP and 56 ns for acrylodan). This suggests that (a) the orientation of the probes is altered with respect to the protein axes and (b) the protein becomes more elongated with an axial ratio of approximately 2.4. For comparison, the hydrodynamic properties of the anthrax AC exotoxin were computed by a mathematical approach (hydropro), which uses the 3D structure [Drum, C.L., Yan, S.-Z., Bard, J., Shen, Y.-Q., Lu, D., Soelalman, S., Grabarek, Z., Bohm, A. & Tang, W.-J. (2002) Nature (London)415, 396-402]. A change in axial ratio is also observed on CaM binding, but in the reverse direction from that for AC: from 1.7 to 1.3. The mechanisms of activation of the two proteins by CaM may therefore be different.  相似文献   

2.
Guha S  Sahu K  Roy D  Mondal SK  Roy S  Bhattacharyya K 《Biochemistry》2005,44(25):8940-8947
Solvation dynamics at the active site of an enzyme, glutaminyl-tRNA synthetase (GlnRS), was studied using a fluorescence probe, acrylodan, site-specifically attached at cysteine residue C229, near the active site. The picosecond time-dependent fluorescence Stokes shift indicates slow solvation dynamics at the active site of the enzyme, in the absence of any substrate. The solvation dynamics becomes still slower when the substrate (glutamine or tRNA(Gln)) binds to the enzyme. A mutant Y211H-GlnRS was constructed in which the glutamine binding site is disrupted. The mutant Y211H-GlnRS labeled at C229 with acrylodan exhibited significantly different solvent relaxation, thus demonstrating that the slow dynamics is indeed associated with the active site. Implications for catalysis and specificity have been discussed.  相似文献   

3.
Equine platelet tropomyosin was labeled with the sulfhydryl-specific fluorescent reagent 6-acryloyl-2-dimethylaminonaphthalene (acrylodan). The extent of labeling at 4 degrees C could be regulated between 0.5 and 1.3 acrylodans per tropomyosin chain by varying the reaction time from 1 to 4.5 h. Acrylodan-labeled platelet tropomyosin, AD-P-TM, was highly fluorescent, having an emission maximum near 518 nm on excitation at 365 nm. Steady-state measurements of polarization of the fluorescence of AD-P-TM in both low and high ionic strength solutions gave Perrin plots that exhibited sharp changes in slope near 50 degrees C, indicative of a sharp increase in mobility of the label at that temperature. This correlates with the melting temperature of the platelet tropomyosin coiled coil observed by circular dichroism [G. P. C?té, W. G. Lewis, M. D. Pato, and L. B. Smillie, (1978) FEBS Lett. 91, 237-241]. Perrin plots of carboxypeptidase A-treated platelet tropomyosin that was labeled with acrylodan after digestion resembled more closely those of acrylodan-labeled cardiac tropomyosin rather than those of AD-P-TM, suggesting that the observed emission arose from label at Cys-153 on each truncated platelet tropomyosin chain. In solutions containing 150 mM KCl and 5 mM MgCl2, addition of actin at up to a sixfold molar excess over AD-P-TM caused both the fluorescence emission intensities and fluorescence polarization values of samples to increase. In the presence of actin, the wavelength of maximal emission was shifted to shorter values by about 5 to 7 nm. These changes indicate that actin does bind to AD-P-TM and that the binding affects the environment of the label, both by making it more hydrophobic and by reducing the freedom of the label to tumble in solution.  相似文献   

4.
Jin S  Kurtz DM  Liu ZJ  Rose J  Wang BC 《Biochemistry》2004,43(11):3204-3213
The X-ray crystal structure of recombinant Desulfovibrio vulgaris rubrerythrin (Rbr) that was subjected to metal constitution first with zinc and then iron, yielding ZnS(4)Rbr, is reported. A [Zn(SCys)(4)] site with no iron and a diiron site with no appreciable zinc in ZnS(4)Rbr were confirmed by analysis of the anomalous scattering data. Partial reduction of the diiron site occurred during the synchrotron X-ray irradiation at 95 K, resulting in two different diiron site structures in the ZnS(4)Rbr crystal. These two structures can be classified as containing mixed-valent Fe1(III)(mu-OH(-))(mu-GluCO(2)(-))(2)Fe2(II) and Fe1(II)(mu-GluCO(2)(-))(2)Fe2(III)-OH(-) cores. The data do not show any evidence for alternative positions of the protein or solvent ligands. The iron and ligand positions of the solvent-bridged site are close to those of the diferric site in all-iron Rbr. The diiron site with only the two carboxylato bridges differs by an approximately 2 A shift in the position of Fe1, which changes from six- to four-coordination. The Fe1- - -Fe2 distance (3.6 A) in this latter site is significantly longer than that of the site with the additional solvent bridge (3.4 A) but significantly shorter than that previously reported for the diferrous site (4.0 A) in all-iron Rbr. The apparent redox-induced movement of Fe1 at 95 K in the ZnS(4)Rbr crystal implies an extremely low activation barrier, which is consistent with the rapid (approximately 30 s(-1)) room temperature turnover of the all-iron Rbr during its catalysis of two-electron reduction of hydrogen peroxide. ZnS(4)Rbr does not show peroxidase activity, presumably because the [Zn(SCys)(4)] site, unlike the [Fe(SCys)(4)] site, cannot mediate electron transfer to the diiron site. One or both of the diiron site structures in the cryoreduced ZnS(4)Rbr crystal are likely to represent that (those) of transient mixed-valent diiron site(s) that must occur upon return of the diferric to the diferrous oxidation level during peroxidase turnover.  相似文献   

5.
Water molecules in hydrophobic biological cleft/cavities are of contemporary interest for the biomolecular structure and molecular recognition of hydrophobic ligands/drugs. Here, we have explored picosecond-resolved solvation dynamics of water molecules and associated polar amino acids in the hydrophobic cleft around Cys-34 position of Endogenous Serum Albumin (ESA). While site selective acrylodan labeling to Cys-34 allows us to probe solvation in the cleft, Förster resonance energy transfer (FRET) from intrinsic fluorescent amino acid Trp 214 to the extrinsic acrylodan probes structural integrity of the protein in our experimental condition. Temperature dependent solvation in the cleft clearly shows that the dynamics follows Arrhenius type behavior up to 60 °C, after which a major structural perturbation of the protein is evident. We have also monitored polarization gated dynamics of the acrylodan probe and FRET from Trp 214 to acrylodan at various temperatures. The dynamical behavior of the immediate environments around the probe acrylodan in the cleft has been compared with a model biomimetic cavity of a reverse micelle (w0 = 5). Using same fluorescent probe of acrylodan, we have checked the structural integrity of the model cavity at various temperatures using picosecond-resolved FRET from Trp to acrylodan in the cavity. We have also estimated possible distribution of donor-acceptor distances in the protein and reverse micelles. Our studies reveal that the energetics of the water molecules in the biological cleft is comparable to that in the model cavity indicating a transition from bound state to quasibound state, closely consistent with a recent MD simulation study.  相似文献   

6.
We searched for new fluorescent probes of catalytic-site nucleotide binding in F(1)F(0)-ATP synthase by introducing Cys mutations at positions in or close to catalytic sites and then reacting Cys-mutant F(1) with thiol-reactive fluorescent probes. Four suitable mutant/probe combinations were identified. beta F410C labeled by 7-fluorobenz-2-oxa-1,3-diazole-4-sulfonamide (ABD-F) gave very large signal changes in response to nucleotide, allowing facile measurement of fluorescence and nucleotide-binding parameters, not only in F(1) but also in F(1)F(0). The results are consistent with the presence of three asymmetric catalytic sites of widely different affinities, with similar properties in both enzymes, and revealed a unique probe environment at the high-affinity site 1. beta Y331C F(1) labeled by ABD-F gave a large signal which monitored catalytic site polarity changes that occur along the ATP hydrolysis pathway. Two other mutant/probe combinations with significant nucleotide-responsive signals were beta Y331C labeled by 5-((((2-iodoacetyl)amino)ethyl)amino)naphthaline-1-sulfonic acid and alpha F291C labeled by 2-4'-(iodoacetamido)anilino)naphthalene-6-sulfonic acid. The signal of the latter responds differentially to nucleoside diphosphate versus triphosphate bound in catalytic sites.  相似文献   

7.
A series of new iron(III) and copper(II) complexes of bovine serum transferrin (BTf), with carbonate and/or oxalate as the synergistic anion, are presented. The complexes [Fe(2)(CO(3))(2)BTf], [Fe(2)(C(2)O(4))(2)BTf], [Cu(2)(CO(3))(2)BTf] and [Cu(C(2)O(4))BTf] were prepared by standard titrimetric techniques. The oxalate derivatives were also obtained from the corresponding carbonate complexes by anion-displacement. The site-preference of the transition metal-oxalate synergism has facilitated the preparation and isolation of the mononuclear complex [Cu(C(2)O(4))BTf], the mixed-anion complexes [Cu(2)(CO(3))(C(2)O(4))BTf] and [Fe(2)(CO(3))(C(2)O(4))BTf] and the mixed-metal complex [FeCu(C(2)O(4))(2)BTf]. The sensitivity of electron paramagnetic resonance (EPR) spectroscopy to the nature of the synergistic anions at the specific-binding sites of the transferrins has made this physical technique particularly indispensable to this study. None of the other members of the transferrin family of proteins has ever been demonstrated to bind the ferric and cupric ions one after the other, each occupying a separate specific-binding site of the same transferrin molecule, as a response to the coordination restrictions imposed by the oxalate ion. The bathochromic shift of the visible p(pi)-d(pi*) CT band for iron(III)-BTf and the hypsochromic shift of the p(pi)-d(sigma*) CT band for copper(II)-BTf, on replacing carbonate by oxalate as the associated anion, are consistent with the relative positions of these anionic ligands in the spectrochemical series and the nature of the d-type acceptor orbitals involved in the CT transitions. The binding and spectroscopic properties of bovine serum transferrin--a serum transferrin--very nearly mirror those of human serum transferrin, but differ significantly from those of human lactoferrin.  相似文献   

8.
Annexin A2 (AnxA2) is a Ca(2+)- and phospholipid-binding protein involved in many cellular regulatory processes. Like other annexins, it is constituted by two domains: a conserved core, containing the Ca(2+) binding sites, and a variable N-terminal segment, containing sites for interactions with other protein partners like S100A10 (p11). A wealth of data exists on the structure and dynamics of the core, but little is known about the N-terminal domain especially in the Ca(2+)-induced membrane-bridging process. To investigate this protein region in the monomeric AnxA2 and in the heterotetramer (AnxA2-p11)(2), the reactive Cys8 residue was specifically labelled with the fluorescent probe acrylodan and the interactions with membranes were studied by steady-state and time-resolved fluorescence. In membrane junctions formed by the (AnxA2-p11)(2) heterotetramer, the flexibility of the N-terminal domain increased as compared to the protein in solution. In "homotypic" membrane junctions formed by monomeric AnxA2, acrylodan moved to a more hydrophobic environment than in the protein in solution and the flexibility of the N-terminal domain also increased. In these junctions, this domain is probably not in close contact with the membrane surface, as suggested by the weak quenching of acrylodan observed with doxyl-PCs, but pairs of N-termini likely interact, as revealed by the excimer-forming probe pyrene-maleimide bound to Cys8. We present a model of monomeric AnxA2 N-terminal domain organization in "homotypic" bridged membranes in the presence of Ca(2+).  相似文献   

9.
The interaction of manganese(II) carboxylate salts [Mn(O(2)CR)(2), where R=ethane, pentane] with H(2)L(1) [N,N'-bis(3-methoxy-2-hydroxybenzaldehyde)-1,2-phenylenediamine] and H(2)L(2) [N,N'-bis(3-ethoxy-2-hydroxybenzaldehyde)-1,2-phenylenediamine] was studied. MnL(1)(O(2)CEt)(H(2)O) (1), MnL(1)(O(2)CPe(n))(H(2)O) (2), MnL(2)(O(2)CEt)(H(2)O)(2) (3) and MnL(2)(O(2)CPe(n))(H(2)O)(2) (4) were isolated and thoroughly characterised by elemental analysis, FAB mass spectrometry, infrared and (1)H NMR spectroscopy, magnetic susceptibility measurements, molar conductivities, and cyclic and normal pulse voltammetry. Compounds 1 and 2 were crystallographically characterised revealing a tetragonally elongated octahedral geometry for the manganese coordination sphere and also a dimeric nature through mu-aqua bridges. Complexes 1-4 behave as efficient peroxidase mimics in the presence of the water-soluble trap ABTS, probably due to their ease to coordinate the substrate molecule. A correlation between rhombicity of the complexes and peroxidase activity has also been established.  相似文献   

10.
We previously reported the construction of a family of reagentless fluorescent biosensor proteins by the structure-based design of conjugation sites for a single, environmentally sensitive small molecule dye, thus providing a mechanism for the transduction of ligand-induced conformational changes into a macroscopic fluorescence observable. Here we investigate the microscopic mechanisms that may be responsible for the macroscopic fluorescent changes in such Fluorescent Allosteric Signal Transduction (FAST) proteins. As case studies, we selected three individual cysteine mutations (F92C, D95C, and S233C) of Escherichia coli maltose binding protein (MBP) covalently labeled with a single small molecule fluorescent probe, N-((2-iodoacetoxy)ethyl)-N-methyl)amino-7-nitrobenz-2-oxa-1,3-diazole (NBD), each giving rise to a robust FAST protein with a distinct maltose-dependent fluorescence response. The fluorescence emission intensity, anisotropy, lifetime, and iodide-dependent fluorescence quenching were determined for each conjugate in the presence and absence of maltose. Structure-derived solvent accessible surface areas of the three FAST proteins are consistent with experimentally observed quenching data. The D95C protein exhibits the largest fluorescence change upon maltose binding. This mutant was selected for further characterization, and residues surrounding the fluorophore coupling site were mutagenized. Analysis of the resulting mutant FAST proteins suggests that specific hydrogen-bonding interactions between the fluorophore molecule and two tyrosine side-chains, Tyr171 and Tyr176, in the open state but not the closed, are responsible for the dramatic fluorescence response of this construct. Taken together these results provide insights that can be used in future design cycles to construct fluorescent biosensors that optimize signaling by engineering specific hydrogen bonds between a fluorophore and protein.  相似文献   

11.
In this work a combined pulsed-laser, time-resolved photoacoustic calorimetry (PAC) and fluorescence study is presented on two widely used covalent protein probes, fluorescein-5-isothiocyanate (FITC) and 6-acryloyl-2-dimethylaminonaphtalene (acrylodan). Three proteins that contain a single free thiol, namely carbonic anhydrase, bovine serum albumin (BSA) and papain, have been selectively labelled with FITC and acrylodan, and their fluorescence emission was quenched with KI. Nonradiative decays of the excited states of FITC are used to complement the information usually obtained by monitoring the quenching of fluorescence emssion. Data analysis evidences the dependence of the nonradiative quenching constants on the exposure of the dye to the solvent, and shows the involvement of a triplet state of FITC in the non radiative deexcitation. The shielding of the binding sites from the solvent is demonstrated also by the fluorescence emission of acrylodan and by the Stern-Volmer analysis of fluorescence quenching by KI. From photoacoustic data, an estimate of the fluorescent quantum yield of bound FITC is obtained. This work demonstrates the complete equivalence of quenching data obtained by fluorescence and photoacoustics measurements and shows that this combined approach allows a better control of the photophysics of the dyes involved in the quenching process.  相似文献   

12.
A simple identification method of L-tyrosine (Tyr) and L-cysteine (Cys) using gemini surfactant coated CdSe quantum dots by using a fluorescent spectroscopic technique is proposed. The gemini surfactant modified QDs show a selective fluorescence response between Tyr and Cys by addition of p-sulfonatocalix[4]arene (pSCA). The CdSe QDs coated with gemini surfactant [C(12)H(25)N(+)(CH(3))(2)(CH(2))(4)(CH(3))(2)N(+)C(12)H(25)].2Br(-) (GS) obviously responds to Tyr. While in the presence of pSCA, it shows selectivity to Cys due to the cooperation of gemini surfactant coated QDs (GS-QDs) and pSCA. Under optimal conditions, it is found that the luminescence of the GS-QDs enhanced by Tyr in a concentration-dependent fashion is described by a Langmuir binding isotherm equation in the range 5 x 10(-8)-10(-5) M. In the presence of pSCA, the luminescence of the GS-QDs enhanced by Cys in a concentration-dependent fashion can also be described by a Langmuir binding isotherm equation in the range 10(-8)-10(-4) M. The possible mechanism is discussed.  相似文献   

13.
We have assigned the biosynthetic processing steps of cathepsin D to intracellular compartments which are involved in its transport to lysosomes in HepG2 cells. Cathepsin D was synthesized as a 51-kDa proenzyme. After formation of 51-55-kDa intermediates due to processing of N-linked oligosaccharides, procathepsin D was proteolytically processed to an intermediate 44-kDa and the mature 31-kDa enzyme. The intersection of the biosynthetic pathway of cathepsin D with the endocytic pathway was labeled with horseradish peroxidase and monitored biochemically by 3,3'-diaminobenzidine cytochemistry. Horseradish peroxidase was used either as a fluid-phase marker to label the entire endocytic pathway or conjugated to transferrin (Tf) to label endosomes only. Directly after biosynthesis cathepsin D was accessible neither to horseradish peroxidase nor Tf-horseradish peroxidase. Newly synthesized 51-55-kDa species of cathepsin D present in the trans-Golgi reticulum were accessible to both horseradish peroxidase and Tf-horseradish peroxidase. The accessibility of trans-Golgi reticulum to both endocytosed horseradish peroxidase and Tf-horseradish peroxidase was monitored by colocalization with a secretory protein, alpha 1anti-trypsin. The proteolytic processing of 51-55-kDa to 44-kDa cathepsin D occurred in compartments which were fully accessible to fluid-phase horseradish peroxidase. Tf-horseradish peroxidase had access to only 20% of 44-kDa cathepsin D while it had no access to 31-kDa cathepsin D. In contrast, the 31-kDa species was completely accessible to fluid-phase horseradish peroxidase. We conclude that proteolytic processing of 51-55-kDa to 44-kDa cathepsin D occurs in endosomes, whereas the processing of 44-31-kDa cathepsin D takes place in lysosomes.  相似文献   

14.
Copper-based transition metal complexes performing single- and double-strand scission of DNA have been studied. The dinuclear complexes [Cu(2)(L)(2)(OCH(3))(2)(NH(3))(2)] and [Cu(2)(L)(2)(OCH(3))(2)(DMSO)(2)] are more active than the corresponding mononuclear [Cu(L)(2)(py)(2)] (where HL= N-(4-methylbenzothiazol-2-yl)benzenesulfonamide), suggesting that the dinuclearity is an important factor in the oxidative cleavage of DNA. The cleavage efficiency of the complexes depends on the reducing agent used in the process, the tandem ascorbate/H(2)O(2) being the most efficient. PAGE analyses have shown that these complexes cleave DNA without sequence selectivity. The DNA degradation process takes place mainly by C1' oxidation, but C4' and C5' oxidations cannot be ruled out as minor pathways. These copper complexes preferably oxidize guanine under mild conditions, but under more drastic conditions the oxidation reactivity appears to be T>G>C>A, suggesting the intervention of hydroxyl radicals as active species.  相似文献   

15.
The protein composition and architecture of the photosynthetic membranes from the cyanobacterium, Synechococcus cedrorum, were analyzed with the aid of site-specific labels. Using membranes labeled with 35S, about 50 membrane proteins can be detected by sodium dodecyl sulfate acrylamide gel electrophoresis. Approximately half of the proteins are accessible to modification by the impermeant probe, lactoperoxidase, indicating that they have surface-exposed domains. At least six of these external proteins can be removed by EDTA washing; the correspondence in molecular weights between five of these EDTA-extractable proteins and those of typical chloroplast coupling factor preparations may indicate that they are subunits of a membrane-bound ATPase. The photoactive, lipophilic compound, [125I]iodonaphthyl azide, was used to label protein domains in contact with the lipid bilayer. Iodonaphthyl azide modification led to a labeling pattern significantly different from that seen with lactoperoxidase. In particular, proteins in the 13 000–20 000 dalton range that were labeled poorly or not at all by lactoperoxidase were heavily modified by iodonaphthyl azide.Photosystem I and II particles, extracted from the membrane by digitonin treatment, were iodinated by lactoperoxidase after isolation. The PS I particles acted as a relatively tight complex, with most of the proteins remaining inaccessible to surface modification. The PS II particles, on the other hand, responded as a more open structure, with most of the subunits yielding to lactoperoxidase iodination. Similar studies on a highly fluorescent, temperature-sensitive mutant of S. cedrorum revealed a different organization of the PS II complex. This mutant, when grown at 40°C, inserts a 51 kdalton polypeptide in place of a 53 kdalton protein. This protein also replaces the 53 kdalton species in the PS II complex of the mutant after 40°C growth. The structure of this complex is altered in that more sites become accessible to lactoperoxidase. This is particularly true of the 51 kdalton protein, which is barely labeled in wild-type PS II complexes.  相似文献   

16.
Studies on diiron dithiolato complexes have proven fruitful for modeling the active site of the [FeFe]-hydrogenases. Here we present a departure from the classical Fe(2)S(2) motif by examining the viability of Fe(2)N(2) butterfly compounds as functional models for the diiron active site of [FeFe]-hydrogenases. Derivatization of Fe(2)(BC)(CO)(6) (1, BC=benzo-[c]-cinnoline) with PMe(3) affords Fe(2)(BC)(CO)(4)(PMe(3))(2), which subsequently undergoes protonation at the Fe-Fe bond. The hydride [(mu-H)Fe(2)(BC)(CO)(4)(PMe(3))(2)]PF(6) was characterized crystallographically as the C(2v) isomer. It represents a rare example of a hydrido diiron complex that exists as observable isomers, depending on the location of the phosphine ligands--diapical and apical-basal. This hydride catalyzes the electrochemical reduction of protons.  相似文献   

17.
We have used a combination of cysteine substitution mutagenesis and site-specific labeling to characterize the structural dynamics of mouse acetylcholinesterase (mAChE). Six cysteine-substituted sites of mAChE (Leu(76), Glu(81), Glu(84), Tyr(124), Ala(262), and His(287)) were labeled with the environmentally sensitive fluorophore, acrylodan, and the kinetics of substrate hydrolysis and inhibitor association were examined along with spectroscopic characteristics of the acrylodan-conjugated, cysteine-substituted enzymes. Residue 262, being well removed from the active center, appears unaffected by inhibitor binding. Following the binding of ligand, hypsochromic shifts in emission of acrylodan at residues 124 and 287, located near the perimeter of the gorge, reflect the exclusion of solvent and a hydrophobic environment created by the associated ligand. By contrast, the bathochromic shifts upon inhibitor binding seen for acrylodan conjugated to three omega loop (Omega loop) residues 76, 81, and 84 reveal that the acrylodan side chains at these positions are displaced from a hydrophobic environment and become exposed to solvent. The magnitude of fluorescence emission shift is largest at position 84 and smallest at position 76, indicating that a concerted movement of residues on the Omega loop accompanies gorge closure upon ligand binding. Acrylodan modification of substituted cysteine at position 84 reduces ligand binding and steady-state kinetic parameters between 1 and 2 orders of magnitude, but a similar substitution at position 81 only minimally alters the kinetics. Thus, combined kinetic and spectroscopic analyses provide strong evidence that conformational changes of the Omega loop accompany ligand binding.  相似文献   

18.
A series of proteins has been examined using time-resolved, pulsed-laser volumetric photoacoustic spectroscopy. Photoacoustic waveforms were collected to measure heat release for calculation of fluorescence quantum yields, and to explore the possibility of photoinduced nonthermal volume changes occurring in these protein samples. The proteins studied were the green fluorescent protein (GFP); intestinal fatty acid binding protein (IFABP), and adipocyte lipid-binding protein (ALBP), each labeled noncovalently with 1-anilinonaphthalene-8-sulfonate (1,8-ANS) and covalently with 6-acryloyl-2-(dimethylamino)naphthalene (acrylodan); and acrylodan-labeled IFABP and ALBP with added oleic acid. Of this group of proteins, only the ALBP labeled with 1,8-ANS showed significant nonthermal volume changes at the beta = 0 temperature (approximately 3.8 degrees C) for the buffer used (10 mM Tris-HCI, pH 7.5) (beta is the thermal cubic volumetric expansion coefficient). For all of the proteins except for acrylodan-labeled IFABP, the fluorescence quantum yields calculated assuming simple energy conservation were anomalously high, i.e., the apparent heat signals were lower than those predicted from independent fluorescence measurements. The consistent anomalies suggest that the low photoacoustic signals may be characteristic of fluorophores buried in proteins, and that photoacoustic signals derive in part from the microenvironment of the absorbing chromophore.  相似文献   

19.
The reaction of trans-[Ru(NH(3))(4)P(OEt)(3)NO](3+) and mitochondria was investigated through differential pulse polarography and fluorimetry. The nitrosyl complex undergoes one-electron reduction centered on the NO ligand site. The reaction between the mitochondrial reductor and trans-[Ru(NH(3))(4)P(OEt)(3)NO](3+) exhibits a second order specific rate constant calculated as k=2 x 10(1) M(-1) s(-1). The reduced species, trans-[Ru(NH(3))(4)P(OEt)(3)NO](2+), quickly releases NO, yielding trans-[Ru(NH(3))(4)P(OEt)(3)H(2)O](2+). The low toxicities of both trans-[Ru(NH(3))(4)P(OEt)(3)(NO)](2+) and trans-[Ru(NH(3))(4)P(OEt)(3)H(2)O](2+) and its ability to release NO after reductive activation in a biological medium make the nitrosyl compound a useful model of a hypotensive drug.  相似文献   

20.
The fluorescent probe, 6-acryloyl-2-dimethylaminonaphtalene (acrylodan) was reacted with the oligomycin-sensitivity conferring protein (OSCP). Acrylodan bound covalently to the single cysteinyl residue of the protein. Acrylodan-OSCP was fully competent in conferring oligomycin sensitivity to the mitochondrial F0-F1 ATPase complex. The fluorescence emission peak of acrylodan-OSCP was blue-shifted compared to that of an acrylodan-mercaptoethanol adduct, which means that acrylodan experiences a hydrophobic environment in OSCP. Binding of acrylodan-OSCP to the isolated F1 was accompanied by a red shift of fluorescence. It was achieved in less than 1 s at 25 degrees C. The titration curve revealed one high affinity OSCP binding site per F1. Acrylodan-OSCP appears to be an interesting tool for studying the dynamics of structural changes within the mitochondrial ATPase complex.  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