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.
Coherent anti-Stokes Raman scattering spectra, in resonance with the isoalloxazine visible electronic transition, have been obtained down to 300 cm?1 for flavin adenine dinucleotide, riboflavin binding protein and glucose oxidase, in H2O and D2O. Several isoalloxazine vibrational modes can be identified by analogy with those of uracil. Of particular interest is a band at ~1255 cm?1 in H2O, which is replaced by another at ~1295 cm?1, in D2O. The H2O band appears to be a sensitive monitor of H-bonding of the N3 isoalloxazine proton to a protein acceptor group. It shifts down by 10 cm?1 in riboflavin binding protein, and disappears altogether in glucose oxidase. Other band shifts, of 3–5 cm?1, are similar for the two flavoproteins, and may reflect environmental changes between aqueous solution and the protein binding pockets.  相似文献   

2.
The dimer formation process of the flavin in aqueous solution has been studied. The difference absorption spectra with the change of concentration in Stokes and anti-Stokes excitation region of the flavomononucleotide and riboflavin were measured. The highest temperature in which the dimers still appear is discussed. It is suggested that this temperature Td can be treated as one of the empirical parameters which describe the dimer formation process of the dyes in solutions. The aqueous solution of flavins with the concentration c?5·10?5 M at room temperature can be treated as a flavin monomers solution. For higher concentrations the flavin monomers and dimers exist in a solution at room temperature.  相似文献   

3.
Riboflavin (RF) plays an important role in various metabolic redox reactions in the form of flavin adenine dinucleotide and flavin mononucleotide. Human serum albumin (HSA) is an important protein involved in the transportation of drugs, hormones, fatty acid and other molecules which determine the biodistribution and physiological fate of these molecules. In this study, we have investigated the interaction of riboflavin RF with HSA under simulative physiological conditions using various biophysical, calorimetric and molecular docking techniques. Results demonstrate the formation of riboflavin–HSA complex with binding constant in the order of 104 M?1. Fluorescence spectroscopy confirms intermediate strength having a static mode of quenching with stoichiometry of 1:1. Experimental results suggest that the binding site of riboflavin mainly resides in sub-domain IIA of HSA and that ligand interaction increases the α-helical content of HSA. These parameters were further verified by isothermal titration calorimetry ITC which confirms the thermodynamic parameters obtained by fluorescence spectroscopy. Molecular docking was employed to suggest a binding model. Based on thermodynamic, spectroscopic and computational observations it can be concluded that HSA-riboflavin complex is mainly stabilized by various non-covalent forces with binding energy of ?7.2 kcal mol?1.  相似文献   

4.
A thermodynamic study of the binding of flavins (FMN, FAD, 8-carboxylic acid-riboflavin) to the purified apoflavodoxin from Azotobacter vinelandii has been conducted. The binding of FMN was studied at a number of temperatures (10,15, 20, 25, and 30 °C), pH's (6.0, 7.4, and 9.0), and buffer conditions. The binding of FAD was studied at pH 7.4 and 25 °C under a number of buffer conditions. The binding of 8-carboxylic acid-riboflavin to the apoflavodoxin and the binding of FMN to the dimeric form of the apoflavodoxin were investigated at pH 7.4 and 25 °C. Enthalpies of binding for FMN, FAD, and 8-carboxylic- acid-riboflavin were ?28.3, ?16.6, and ?14.0 kcal mol?1, respectively. The enthalpy of binding of FMN to the dimeric form of the apoflavodoxin was ?22.2 kcal mol of binding sites?1. Binding constants of about 108,106, and 106 were obtained for the binding of FMN, FAD, and 8-carboxylic acid-riboflavin, respectively. Using established thermodynamic relationships free energy and entropy changes were calculated. The entropy data indicate that a large degree of ordering of the system occurs upon flavin binding. The pH data suggest that FMN may bind in both the mono-and dianion forms, and that binding doesn't change the pKa of any functional group in the system. It appears that the phosphate group is probably responsible for approximately half the binding enthalpy observed for the binding of FMN. The temperature-dependence data over the temperature range studied is biphasic, centered at 20 °C, indicating that flavin binding occurs to the protein in two thermodynamic states corresponding to the two heat capacities observed. These findings are used to discuss a model for flavin binding.  相似文献   

5.
Ulrike Dohrmann 《Planta》1983,159(4):357-365
Several types of membrane-localized flavin binding sites were investigated in sporangiophores (spph) and mycelia of Phycomyces blakesleeanus. In-vitro binding of riboflavin, riboflavin-5′-phosphate, and flavin-adenine-dinucleotide was demonstrated with unfractionated membrane preparations by means of competition of [14C]riboflavin binding. Saturation of binding was only obtained with the highly water-soluble riboflavin-5′-phosphate, but by extrapolation it was shown that riboflavin showed the highest affinity towards the binding sites (KD about 4·10-6M). The number of binding sites was estimated to be 0.7 nmol g-1 fresh-weight equivalent. Analysis of endogenous soluble flavin revealed that only riboflavin, riboflavin-5′-phosphate, and flavin-adenine-dinucleotide occurred in Phycomyces, and at a concentration of at least 1 nmol g-1 fresh-weight equivalent in entire spph. Thus, the measured binding sites could reach saturation in-vivo. In the apical part of spph to which blue-light sensitivity is restricted, the amount of soluble flavin was three-fold higher. Exclusively in this zone, heat-labile riboflavin proteins were measured at a concentration of about 3 nmol g-1 fresh-weight equivalent. The amount of covalently bound flavin was higher in spph tips than in intact spph (8 nmol and 3 nmol g-1 fresh-weight equivalent, respectively). In either case, the concentrations of the flavin-membrane complexes were higher than the theoretical calculated concentration of (anisotropic) blue-light photoreceptor in Phycomyces (Bergman et al. 1969), and their involvement in blue-light photoreception is considered.  相似文献   

6.
The properties of the riboflavin-binding site in the riboflavin-binding protein from egg white have been elucidated by determining constants for binding of flavin analogs to the protein and by changes in absorption spectra of free and bound flavins. The spectral changes and unfavorable interaction of the protein with charged species indicate that the overall flavin environment in the holoprotein is hydrophobic. Modification of either ring or side-chain portions of flavin usually results in a decrease of binding energy. Although no one portion of the structure is absolutely essential, both 7- and 8-methyl groups and 2′-hydroxyl group contribute most significantly to binding. The binding site at the region of C-2 and N-3 of the isoalloxazine is rather insensitive to the relative site of a substituent and thus relatively open, whereas considerable steric limitation is imposed at C-8, N-10, especially C-1′, and 4carbonyl positions. The hydroxyl groups of the N-10 side chain contribute in a stereoselective manner by formation of hydrogen bonds. Studies with model compounds that represent only a part of flavin suggest that the dimethylbenzenoid portion of the ring is involved in primary interactions of binding, and relatively buried in the protein. The quenching of protein fluorescence upon binding is mainly due to ground-state stacking interaction between a trytophanyl residue at the binding site and the quinoxaline portion, and not to Förster energy transfer.  相似文献   

7.
Resonance Raman (RR) spectra excited at 632.8 nm within a charge transfer absorption band were obtained for a catalytic intermediate, the purple complex of D-amino acid oxidase with D-proline or D-alanine as a substrate. The resonance enhanced Raman lines around 1605 and 1360 cm?1 in either of the complexes were suggested to be derived from vibrational modes of reduced flavin molecule. Since the highest energy band at 1692 cm?1 in the RR spectrum with D-alanine was shifted to 1675 cm?1 upon [15N] substitution of alanine and ammonium, this Raman line in the spectrum with D-alanine or the line at 1658 cm?1 with D-proline is assigned to the CN stretching mode of an imino acid corresponding to each amino acid. These results confirm the concept that the purple intermediate of D-amino acid oxidase consists of reduced flavin and an imino acid.  相似文献   

8.
In order to analyze the melting behavior of 5′-rGMP gel at acidic pH and self-aggregate near neutral pH we have obtained Raman spectra of aqueous solutions of 5′-rGMP at various temperatures. At low temperature the intensities of Raman peaks at 502, 585, 1083, 1179, 1322, 1366, 1487, and 1578 cm?1 decrease due to the formation of ordered structure (Raman hypochromism). In contrast, the peaks at 671, 725, 813, and 1338 cm?1 become stronger at low temperature (Raman hyperchromism). The Raman hyperchromism of the 671- and 813-cm?1 peaks have been explained in terms of detailed structural models. Recently, the 668- and 682-cm?1 peaks in the Raman spectrum of aqueous 5′-rGMP solution have been attributed to the guanine ring breathing vibrations in C3′- and C2′-endo conformers [Benevides, J. B., Lemur, D. & Thomas, G. J., Jr. (1984) Biopolymers 23 , 1011–1024]. On the basis of this information our Raman data can be interpreted to suggest that the continuous helix model of 5′-rGMP gel is right-handed. The 1487-cm?1 peak intensity has been used to monitor the melting profies at several pHs. Near neutral pH the melting profile shows a single transition, whereas at acidic pH it shows two transitions. From these observations we propose possible pathways for the melting of 5′-rGMP gel formed at acidic pH and self-aggregate formed near neutral pH.  相似文献   

9.
The Raman spectra of Bence-Jones proteins (BJP) were measured for their native and denatured states. All of the native BJPs investigated gave amide I at 1670–1675 cm?1 and amide III at 1242–1246 cm?1. Although the amide I was shifted to 1667 cm?1 upon the LiBr, acid, and thermal denaturation, as expected, the amide III frequency was unaltered, indicating that the antiparallel β- and disordered structures of BJP provide amide III at almost the same frequencies. The intensity of the 880-cm?1 line of native BJP was relatively intense compared with that of amino acid mixed solution in which the mole ratios of Trp, Phe, and Tyr were adjusted to reproduce the corresponding ratios of BJP. However, the intensity was evidently reduced upon LiBr, acid, and thermal denaturation, approaching that of the amino acid mixture. Thus, the intensity of the 880-cm?1 line is proposed as a practical probe for the environment of Trp residues. The pH dependence of the intensity of the 880-cm?1 line suggests that one of two buried Trp residues is exposed between pH 4 and 3.2 and the other between pH 3.2 and 1.4. The variable fragment (VL) of BJP (Tod) exhibited a S? S stretching Raman line at 525 cm?1. Provided that the crystallographic data of the VL of BJP is applicable to VL of BJP (Tod), the 525 cm?1 of the S? S stretching frequency should be assigned to a TGG conformation of linkage, but not to the AGT or AGG conformation. This supports Sugeta's model rather than Scheraga's model.  相似文献   

10.
11.
A riboflavin-binding hexamerin isolated from pupal hemolymph of Hyalophora cecropia has a native Mr of 510,000, subunit Mr of 85,000, and a 5% carbohydrate content. An intrachain cross-link was confirmed in protease limit digests. Ellman titration confirmed the presence of a sulfhydryl group, which is needed for this linkage. Though Cu2+ is known to promote the linkage, heavy metals were not detected in the isolate. Heat denaturation released ligand with the absorbency, fluorescence spectra, and chromatographic behavior of riboflavin. Binding resulted in substantial quenching of the fluorescence of both the isoalloxazine in riboflavin and of aromatic groups in the apoprotein. Kinetic analysis indicated a KD of 2.5 × 10?7 M for riboflavin, 1.3 × 10?7 M for lumiflavin, and greater than 1 × 10?6 M for FMN and FAD. Over four moles of flavin were bound per mole of hexamerin. The amount of riboflavin in pupal hemolymph is sufficient to occupy only 2–3 of these sites. Riboflavin is also associated with lipophorin and vitellogenin, but the molar ratios after protein isolation were low. On a standard laboratory diet, riboflavin is in great excess, but most of it is apparently excreted before the apoprotein first appears in the hemolymph, just before wandering. The concentration of riboflavin-binding hexamerin rises to 15–30 mg/ml in pupae; relative to other hexamerins, very little is stored in the fat body. All of the apoprotein and 75% of riboflavin disappear from the hemolymph during adult development. An amount of flavin at least equal to that stored in pupal hemolymph is transferred to the eggs formed during this period. © 1994 Wiley-Liss, Inc.  相似文献   

12.
The interactions of methylcobalamin (CH3-B12) with Pt(CN)42?, PtCl42?, and Pt(SCN)42? in aqueous solution were studied by UV-visible and 1H NMR spectroscopy. Together with earlier results on the mechanism of the Pt(IV)-dependent methyl-transfer reaction from CH3-B12 to Pt(II), these studies suggest at least three Pt binding sites on CH3-B12. One site, which is occupied by all three complexes (K1 = 4 X 103 M?1 for Pt(CN)42? and 3 X 103 M?1 for PtCl42?), is located on the CoCH3 side of the corrin macrocycle, and is involved in the methyl-transfer process in the presence of a Pt(IV) complex. An additional site for Pt(SCN)42? is the N-3 of the benzimidazole group, resulting in dissociation of this group from the cobalt. An additional site for Pt(CN)42? has a binding constant of 16 M1? and 1H NMR changes indicate perturbation but not dissociation of the benzimidazole group. Only the first interaction is discerned for PtCl42?.  相似文献   

13.
Dihydroxyanthraquinone (DHAQ) and ametantrone (anthraquinone) are two new anthracenedione antineoplastic agents which were found by proton NMR spectroscopy to self-associate in aqueous media. Self-association was consistent with a bimolecular model, with average association constant values of 3400 and 2900 m?1 determined for DHAQ and ametantrone, respectively. Both anthracenediones interacted with the flavin nucleotides FMN and FAD to produce concentration-dependent upfield shifts of the flavin isoalloxazine ring proton signals, as observed by proton NMR spectroscopy. Average association constant values obtained for FMN-DHAQ, FAD-DHAQ, FMN-metantrone, and FAD-ametantrone complexation were 5100, 2600, 4300, and 1600 m?1, respectively. Optical difference spectroscopy confirmed FMN-DHAQ complexation, which resulted in a hyperchromic, bathochromic shift of the DHAQ spectrum following addition of FMN. These results were consistent with the formation of a ππ bimolecular ring-stacking complex. Information obtained on anthracenedione self-association and complexation with flavins may be of consequence in the interpretation of anthracenedione-DNA binding data and flavoprotein-mediated anthracenedione metabolic activation.  相似文献   

14.
S P Fodor  P A Starr  T G Spiro 《Biopolymers》1985,24(8):1493-1500
Raman spectra have been recorded for poly(dG-dT) · poly(dA-dC) and poly(dA-dT) · poly(dA-dT) in low salt and at high concentrations of CsF. Poly(dG-dT) · poly(dA-dC) shows no change in the 682-cm?1 guanine mode, demonstrating the absence of the Z-structure at high salt. The 790-cm?1 phosphodiester symmetric stretch, however, shifts up 5 cm?1 in 4.3M CsF, suggesting a slight conformational change, associated with ion binding or hydration changes. Poly(dA-dT) · poly(dA-dT) shows an additional broad band at 816 cm?1, attributed to the phosphodiester modes associated with the C3′-endo deoxyribose units in the alternating B-structure. In this case, both the 841- and the 816-cm?1 asymmetric phosphodiester stretches, associated with the C2′- and C3′-endo units, shift down on addition of CsF in a sequential manner. Correlation of this sequence with that previously observed for the two 31P-nmr resonances, establishes that the phosphodiester stretching frequencies depend on the conformation of the 5′-sugar, and not on the 3′-sugar.  相似文献   

15.
Both extensive theoretical calculations and experimental data obtained during several decades leave little doubt that flavin adenine dinucleotide (FAD) exists in an open as well as in a closed conformation in aqueous solution. However, the knowledge about the intramolecularly stacked complex of FAD is constructed on indirect methods while direct structural evidence is lacking. Recently, dodecin was reported as an unspecific flavin binding protein which exhibits the unique binding mode of incorporating stacked dimers of flavins into a single binding pocket. Here, we show that FAD is not bound in this manner, but in monomers of intramolecularly stacked conformation. As resulting from the dodecin ligand binding characteristic, this FAD stacked conformation suggests to be directly sequestered from the aqueous solution and thus to be the first X-ray structural view on a FAD solution-stacked form. Moreover, in extraordinary FAD binding, dodecin serves as a model for studying bound monomeric (FAD) versus bound dimeric (e.g. riboflavin) flavin properties.  相似文献   

16.
Stopped-flow studies on calcium binding to calmodulin showed that under pseudo first order conditions the reaction was complete within 2.5 milliseconds. The time course for calcium dissociation from the native protein showed a single kinetic phase (τ1?1 = 10S?1) while that from the dansylated derivative revealed a second slower kinetic phase (τ1?1 = 10S?1, τ2?1 = 0.31S?1) that accounted for about one-half of the total fluorescence decrease. Therefore the dansyl derivative of calmodulin may provide a useful tool for studying conformational changes in the protein not reflected by the active site tyrosines.  相似文献   

17.
The Raman spectra of β-lactoglobulin in the crystalline, freeze-dried, and solution states are compared. The spectra of the freeze-dried and crystalline proteins were practically identical. The conformationally sensitive amide III line appearing at 1242 cm?1 increased in intensity 30% upon dissolution of the protein in water which is interpreted as a conformational change in the disordered chains of the protein. This result appears to be a phenomenon for globular proteins containing a large disordered chain fraction. The alkaline denaturation of β-lactoglobulin was studied. When the pH was increased from 6.0 to 11.0, the amide III line shifted from 1242 to 1246 cm?1, broadened, and decreased in intensity. This is consistent with the conversion of β-sheet regions in β-lactoglobulin to the disordered conformation, as has been proposed by other investigators. At pH 13.5 the amide III shifts to 1257 cm?1 characteristic of a completely disordered protein, indicating that any remaining “core” of β-sheet has been randomized. Several changes in the intensities of the tyrosine and tryptophan vibrations accompany the denaturation. As the pH is increased from 6.0 (native state) to 11.0 (denatured state) the intensity ratio of two tyrosine ring vibrations, I855 cm?1/I830 cm?1, decreases from 1.0:0.9 to 1.0:1.3. The same ratio for a copolymer consisting of 95% glutamic acid and 5% tyrosine at pH 7.0, where the polymer forms a random coil exposing the tyrosine to the aqueous environment, is 1.0:0.62. This ratio more closely resembles that corresponding to β-lactoglobulin at pH 6.0 (native state) than pH 11.0 (denatured state) suggesting that the average tyrosine in the denatured state may be in a more hydrophobic environment than in the native state. A time-dependent polymerization of the denatured protein reported by other investigators and observed by us may account for the change in the tyrosine environment. A tryptophan vibration appearing at 833 cm?1 in the spectrum of the native state becomes weak as the pH is increased to 11.0. The intensity of this line may also reflect the local environment of the tryptophan residue.  相似文献   

18.
A competitive binding procedure that can be used to determine either riboflavin or riboflavin-binding protein has been developed. Riboflavin-binding protein from chicken egg white binds tightly to DEAE-cellulose while free riboflavin does not. Stock [2-14C]riboflavin solutions, diluted with varying amounts of a standard unlabeled riboflavin solution or an unknown sample, are mixed with aporiboflavin-binding protein and washed through small DEAE-cellulose columns. The protein-bound riboflavin is batch eluted into scintillation vials, counted, and the unknown samples compared to a standard curve. This is a simple, rapid method for assaying riboflavin by isotope dilution. By a slight modification of the incubation conditions of this procedure, the degree of saturation and amount of riboflavin-binding protein can be determined. Data from both assays can be represented by linear plots in which slopes or intercepts correspond to unknown values. The principles presented here have been extended to the assay of biotin and avidin and should apply to other vitamins and vitamin-binding proteins.  相似文献   

19.
Representative examples of the various classes of flavoproteins have been converted to their apoprotein forms and the native flavin replaced by 8-mercapto-FMN or 8-mercapto-FAD. The spectral and catalytic properties of the modified enzymes are characteristically different from one group to another; the results suggest that flavin interactions at positions N(1) or N(5) of the flavin chromophore have profound influences on the properties of the flavoprotein. 1. The 8-thiolate anion form of 8-mercaptoflavin has an absorption maximum in the region 520 to 550 nm epsilon approximately 30 mM-1 cm-1). This form is retained on binding to flavoproteins whose physiological reactions involve obligatory one-electron transfers (e.g. flavodoxin, NADPH-cytochrome P-450 reductase). In the native form these enzymes stabilize the blue neutral radical of the flavin. A radical form of 8-mercaptoflavin is also stabilized by these proteins. 2. The p-quinoid form of 8-mercaptoflavin has an absorption maximum in the range 560 to 600 nm (epsilon approximately 30 mM-1 cm-1). This form is stabilized on binding to flavoproteins of the dehydrogenase-oxidase class (e.g. glucose oxidase, D-amino acid oxidase, lactate oxidase, Old Yellow Enzyme). These same enzymes in their native flavin form stabilize the red semiquinone, and have a pronounced reactivity with sulfite to form flavin N(5)-sulfite adducts. These properties of the native enzyme, including the ability to react with nitroalkane carbanions, are not exhibited by the 8-mercaptoflavoproteins. 3. A group of flavoenzymes fails to conform strictly to the above classification, exhibiting some properties of both classes. These include the examples of flavoprotein hydroxylases and transhydrogenases studied. 4. The riboflavin-binding protein of hen egg whites binds 8-mercaptoriboflavin preferentially in the unionized state, resulting in a shift in pK from 3.8 with free 8-mercaptoriboflavin to greater than or equal to 9.0 with the protein-bound form.  相似文献   

20.
Systematic thermodynamic studies have been conducted for flavin (FMN, FAD) binding to purified riboflavin-binding proteins from hen egg white and egg yolk. These studies were conducted under a variety of temperature (14, 26, and 38 °C), pH (4.5, 5.5, 6.5, 7.4, and 9.0), and buffer conditions, and an extensive thermodynamic profile was constructed. Enthalpies of binding FMN to white riboflavin-binding protein and yolk riboflavin-binding protein were ?19.3 and ?14.4 kcal/mol, respectively, at pH 7.4 and 38 °C. FAD bound to white and yolk riboflavin-binding proteins under the same conditions with ΔH values of ?11.7 and ?6.0, respectively. Binding constants of about 105 and 104 were obtained for FMN and FAD, respectively, and were the same for both proteins under all conditions studied. Using established thermodynamic relationships, we were able to calculate entropy and free energy changes. Entropies indicated a large degree of ordering in the system upon flavin binding with FMN (about ?40 cal/mol/ °C) twice as large as FAD (about ?15 to ?25 cal/mol/ °C), which may indicate a structured solvent interaction with the charged phosphate group, or steric limitations placed on the ribityl side chain in the bound state. Our thermodynamic data support the idea that flavin binding is a mixture of forces, with no one predominant. Analysis of the data suggests that the nucleotide may bind both as the mono- or dianion, that flavin binding occurs with no significant change in the pK of any functional group in the system, except at low pH for FAD binding, and that the temperature variation of the enthalpy change is quite small. These findings are combined with other published data to outline a general scheme of flavin binding with a histidine residue implicated in hydrogen bonding to the adenine portion of FAD, which may be in the unstacked form.  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