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.
Nicotinoprotein alcohol dehydrogenases are enzymes that contain non-dissociable NAD(P)(H) in the active site. The suitability of a nicotinoprotein alcohol dehydrogenase as coenzyme-independent alternative to classic alcohol dehydrogenases for enantioselective synthetic applications was studied. To this end the NADH-containing nicotinoprotein, np-ADH, from Rhodococcus erythropolis DSM 1069 was used as a model enzyme in different types of conversion: asymmetric synthesis, kinetic resolution and racemization. The enzyme was found to catalyze the asymmetric reduction of ketones using cheap reductants, such as ethanol, with high stereoselectivity, but the reaction was too slow to obtain good yields. Kinetic resolutions of racemic alcohols failed due to dismutation of the aldehyde that was used as cosubstrate. Racemization of a secondary alcohol via the corresponding ketone could not be achieved, which was due to an unidentified side reaction. This evaluation shows that, for developing biotransformations of industrial interest using nicotinoprotein alcohol dehydrogenases, the attention should be focused on enzymes with a higher reactivity towards prochiral ketones and secondary alcohols.  相似文献   

2.
The Bacillus methanolicus methanol dehydrogenase (MDH) is a decameric nicotinoprotein alcohol dehydrogenase (family III) with one Zn(2+) ion, one or two Mg(2+) ions, and a tightly bound cofactor NAD(H) per subunit. The Mg(2+) ions are essential for binding of cofactor NAD(H) in MDH. A B. methanolicus activator protein strongly stimulates the relatively low coenzyme NAD(+)-dependent MDH activity, involving hydrolytic removal of the NMN(H) moiety of cofactor NAD(H) (Kloosterman, H., Vrijbloed, J. W., and Dijkhuizen, L. (2002) J. Biol. Chem. 277, 34785-34792). Members of family III of NAD(P)-dependent alcohol dehydrogenases contain three unique, conserved sequence motifs (domains A, B, and C). Domain C is thought to be involved in metal binding, whereas the functions of domains A and B are still unknown. This paper provides evidence that domain A constitutes (part of) a new magnesium-dependent NAD(P)(H)-binding domain. Site-directed mutants D100N and K103R lacked (most of the) bound cofactor NAD(H) and had lost all coenzyme NAD(+)-dependent MDH activity. Also mutants G95A and S97G were both impaired in cofactor NAD(H) binding but retained coenzyme NAD(+)-dependent MDH activity. Mutant G95A displayed a rather low MDH activity, whereas mutant S97G was insensitive to activator protein but displayed "fully activated" MDH reaction rates. The various roles of these amino acid residues in coenzyme and/or cofactor NAD(H) binding in MDH are discussed.  相似文献   

3.
Analysis at 0.25 nm resolution of the crystal structures of lysozyme-Gd(III) and lysozyme-Gd(III)-N-acetyl-D-glucosamine (GlcNac), prepared by diffusion methods, show that there are two main binding positions for Gd(III), one of which is close to glutamic acid-35 and the other close to aspartic acid-52. The two sites are 0.36 nm part. There is no evidence for the weak binding of Gd(III) to any of the eight other carboxy groups of lysozyme. In the presence of Gd(III), the binding of GlcNac is similar to that observed for the binding of the beta-anomer in subsite C. There are numerous small conformational changes in the protein on binding (Gd(III) and the sugar, and these have been quantified to a first approximation by real-space refinement. These changes are similar in both structures, and involve, among other small movements, shifts of one of the disulphide bridges by up to 0.05 nm. The movement of residues 70--74 observed in the binary complex of lysozyme-GlcNac [Perkins, Johnson, Machin & Phillips (1978) Biochem. J. 173-617] is not observed in the ternary complex of lysozyme-Gd(III)-GlcNac. The nature of the lysozyme-Gd(III) complex is discussed in the light of evidence from other crystallographic studies and n.m.r. solution studies. Preliminary findings for a lysozyme-Gd(III) complex prepared by co-crystallization methods are reported.  相似文献   

4.
The active site metal in horse liver alcohol dehydrogenase has been studied by metal-directed affinity labeling of the native zinc(II) enzyme and that substituted with cobalt(II) or cadmium(II). Reversible binding of bromoimidazolyl propionic acid to the cobalt enzyme blueshifts the visible absorption band originating from the catalytic cobalt atom at 655 to 630 nm. Binding of imidazole to the cobalt(II) enzyme redshifts the 655 nm band to 667 nm. Addition of bromoimidazolyl propionic acid blueshifts this 667 nm band back to 630 nm. This proves direct binding of the label to the active site metal in competition with imidazole. The affinity of the label for the reversible binding site in the three enzymes follows the order Zn ? Cd ? Co. After reversible complex formation, bromoimidazolyl propionic acid alkylates cysteine-46, one of the protein ligands to the active site metal. The nucleophilic reactivity of this metal-mercaptide bond in each reversible complex follows the order Co ? Zn ? Cd.  相似文献   

5.
Interactions of stromelysin with a series of inhibitors representative of three chemical templates with distinct binding modes were examined. Unfolding temperatures for inhibitor complexes were 10 degrees C to 15 degrees C greater than for apo stromelysin. Minor changes in ellipticity in the far-UV CD spectra of complexes indicated that ligand-induced conformational changes were localized to the binding site and did not involve gross changes in protein folding. Isothermal titrating calorimetry of thiadiazole-containing inhibitors, which bind in the S(1)-S(3) subsites of stromelysin, indicated that the binding interaction was exothermic and only slightly favorable entropically. Near-UV CD spectra showed large positive ellipticity increases from 250 to 300 nm, consistent with an interaction between the benzene ring of the inhibitor and stromelysin residues Tyr155 and Tyr168. Interactions between stromelysin and amide-hydroxamate ligands, which bind in the S(')(1)-S(')(3) subsites, were found to be both enthalpically and entropically driven. Binding of this class of ligands resulted in modest negative ellipticity changes at 260-285 nm and positive increases at 292 nm. Stromelysin complexed to a lactam-hydroxamate inhibitor with structure extending into both the S(1)-S(3) and S(')(1)-S(')(3) subsites showed increased ellipticity at 245 nm and negative changes at 260-285 and 295 nm.  相似文献   

6.
(-)-Epigallocatechin 3-O-gallate (EGCG) a molecule found in green tea and known for a plethora of bioactive properties is an inhibitor of heat shock protein 90 (HSP90), a protein of interest as a target for cancer and neuroprotection. Determination of the spectral properties of EGCG fluorescence in environments similar to those of binding sites found in proteins provides an important tool to directly study protein-EGCG interactions. The goal of this study is to examine the spectral properties of EGCG fluorescence in an aqueous buffer (AB) at pH=7.0, acetonitrile (AN) (a polar aprotic solvent), dimethylsulfoxide (DMSO) (a polar aprotic solvent), and ethanol (EtOH) (a polar protic solvent). We demonstrate that EGCG is a highly fluorescent molecule when excited at approximately 275 nm with emission maxima between 350 and 400 nm depending on solvent. Another smaller excitation peak was found when EGCG is excited at approximately 235 nm with maximum emission between 340 and 400 nm. We found that the fluorescence intensity (FI) of EGCG in AB at pH=7.0 is significantly quenched, and that it is about 85 times higher in an aprotic solvent DMSO. The Stokes shifts of EGCG fluorescence were determined by solvent polarity. In addition, while the emission maxima of EGCG fluorescence in AB, DMSO, and EtOH follow the Lippert-Mataga equation, its fluorescence in AN points to non-specific solvent effects on EGCG fluorescence. We conclude that significant solvent-dependent changes in both fluorescence intensity and fluorescence emission shifts can be effectively used to distinguish EGCG in aqueous solutions from EGCG in environments of different polarity, and, thus, can be used to study specific EGCG binding to protein binding sites where the environment is often different from aqueous in terms of polarity.  相似文献   

7.
When the benzo(a)pyrene diol epoxide (±)-7β,8α-dihydroxy-9α,10α-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene (BPDE) is mixed into a DNA solution, a 10nm red shift in the absorption maximum of BPDE appears at 354nm which is due to a non-covalent intercalation complex. The major reaction pathway at this intercalation site is the hydrolysis of BPDE to its tetraol which is accompanied by a decrease in the absorbance and a shift from 354 to 353nm (the latter is due to intercalated tetraol). The non-covalent binding constants are approximately 8200M?1 for BPDE and 3300M?1 for the tetraol at 25°C, pH 7.0. Covalent adduct formation between BPDE and DNA occurs either at another, external binding site, or after some rearrangement of the intercalated BPDE, since covalent adducts display a 345nm absorption maximum (2nm red shift only).  相似文献   

8.
The binding between thermolysin and its specific inhibitor, talopeptin (MKI), was found to show a fluorescence increase when excited at 280 nm and 295 nm, and a difference spectrum characterized by two peaks at 294 nm and 285 nm with a shoulder around 278 nm, indicating a microenvironmental change in tryptophan residue(s) of thermolysin and/or talopeptin. The inhibitor constant of talopeptin against thermolysin, Ki, was determined over the pH range 5-9 from the inhibition of the enzyme activity towards 3-(2-furylacryloyl)-glycyl-L-leucine amide (FAGLA) as a substrate. The dissociation constant of thermolysin-talopeptin complex, Kd, determined directly from fluorometric titration was in good agreement with the inhibitor constant, Ki, between pH 6 and 8.5. The pH dependence of Ki and Kd suggested that at least two ionizable groups of thermolysin in their protonated forms are essential for the binding between thermolysin and talopeptin. The temperature dependence of K1 at pH 5.5 indicated that the binding is largely exothermic (delta H degree = -12 kcal/mol) and essentially enthalpy-driven.  相似文献   

9.
Ternary copper(II) complex [Cu(TAN)(O2CMe)] (1), where H-TAN is 1-(2-thiazolylazo)-2-naphthol, is prepared and structurally characterized by X-ray crystallography. The complex has a distorted square pyramidal (4+1) CuN2O3 coordination geometry with the acetate showing chelating axial-equatorial binding mode and TAN as a tridentate ligand bonded to the metal in the basal plane. Complex 1 is one-electron paramagnetic and displays ligand-to-metal charge transfer bands at 575 and 398 nm in dimethylformamide. The reactions of 1 with bases (B) like 1,10-phenanthroline (phen) and kanamycin-A (kan-A) afford ternary complexes of formulation [Cu(TAN)B]+ (B=phen, 2; kan-A, 3) under in situ reaction conditions. Complexes 2 and 3, prepared to explore their DNA binding and photo-induced DNA cleavage activity, display good binding propensity to calf thymus (CT) DNA giving a relative order: 2-3>1. The apparent binding constant (Kapp) for 1 is determined as 9.8 x 10(5)M(-1) from fluorescence quenching experiments using ethidium bromide. The quenching constants (K) values of 1-3, obtained from the Stern-Volmer plots, are 0.28, 0.52 and 0.49, respectively. All the complexes show photo-induced DNA cleavage activity when irradiated with a monochromatic UV light of 365 nm wavelength. A 200 microM complex 1 cleaves approximately 75% supercoiled (SC) DNA on 2h exposure time at 365 nm. A 50 microM solution of 1 in presence of 100 microM phen and kanamycin-A cleaves approximately 99% and approximately 60% SC DNA to its nicked circular form, respectively, for an exposure of 30 min. The complexes also exhibit significant cleavage of SC DNA on irradiation with visible light of wavelengths 532, 575 and 632.8 nm. Control experiments reveal the minor groove binding nature of the complexes. The cleavage reactions involve the formation reactive hydroxyl species as significant inhibition in the presence of dimethyl sulfoxide (DMSO) and catalase is observed. There is no apparent inhibition in cleavage in the presence of singlet oxygen quenchers like sodium azide. The cleavage activity has been found to be higher at the CT band position of 575 nm in comparison to those at 532 and 632.8 nm. The results indicate the involvement of the CT band in the photo-excitation process.  相似文献   

10.
Using CD measurements we show that the interaction of netropsin to poly(dA-dT).poly(dA-dT) involves two binding modes at low ionic strength. The first and second binding modes are distinguished by a defined shift of the CD maximum and the presence of characteristic isodichroic points in the long wavelength range from 313 nm to 325 nm. The first binding mode is independent of ionic strength and is primarily determined by specific interaction to dA.dT base pairs. Employing a netropsin derivative and different salt conditions it is demonstrated that ionic contacts are essential for the second binding mode. Other alternating duplexes and natural DNA also exhibit more or less a second step in the interaction with netropsin observable at high ratio of ligand per nucleotide. The second binding mode is absent for poly(dA).poly(dT). The presence of a two-step binding mechanism is also demonstrated in the complex formation of poly(dA-dT).poly(dA-dT) with the distamycin analog consisting of pentamethylpyrrolecarboxamide. While the binding mode I of netropsin is identical with its localization in the minor groove, for binding mode II we consider two alternative interpretations.  相似文献   

11.
Mercuric binding studies at pH 10 revealed that poly(dA): poly(dT) exhibits a more dramatic absorption spectral alteration than the alternating polymer poly(dA-dT):poly(dA-dT) and induces a unique intense positive CD band at 296 nm during the spectral titrations. Comparative studies with its component single strands suggest that the spectral alterations exhibited by poly(dA): poly(dT) are consistent with a binding model in which the mercuric ions initially bind to thymines and cause the eventual strand separation of the duplex, with subsequent high cooperative binding to the poly(dA) strands. This interpretation is supported by the binding isotherms indicating much stronger mercuric binding to poly(dT) than to poly(dA), with saturation binding densities of 1 Hg(II) per 2 bases and 1 Hg(II) per base, respectively, and very high binding cooperativity for poly(dA). Striking spectral alterations are exhibited by the mercuric binding to poly(dA), likely the consequence of binding to the amino group of dA in an alkaline solution. The mononucleoside dA exhibits minor spectral alterations upon similar mercuric chloride additions whereas the dinucleoside monophosphate d(AA) exhibits significant spectral changes, albeit less pronounced than those of poly(dA). Some sequence effects on the mercuric binding are observed in the dinucleotide studies. Our CD results on the mercuric binding to polynucleotides do not support the contention of (psi)-type condensed complex formation.  相似文献   

12.
The interaction mechanism of lipopeptide antibiotic daptomycin and polyamidoamine (PAMAM) dendrimers was studied using fluorescence spectroscopy. The fluorescence changes observed are associated with daptomycin–dendrimer interactions. The binding isotherms were constructed by plotting the fluorescence difference at 460 nm from kynurenine (Kyn‐13) of daptomycin in the presence and absence of dendrimer. A one‐site and two‐site binding model were quantitatively generated to estimate binding capacity and affinity constants from the isotherms. The shape of the binding isotherm and the dependence of the estimated capacity constants on dendrimer sizes and solvent pH values provide meaningful insight into the mechanism of interactions. A one‐site binding model adequately describes the binding isotherm obtained under a variety of experimental conditions with dendrimers of various sizes in the optimal binding pH region 3.5 to 4.5. Comparing the pH‐dependent binding capacity with the ionization profiles of daptomycin and dendrimer, the ionized aspartic acid residue (Asp‐9) of daptomycin primarily interact with PAMAM cationic surface amine. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
Three distinctive heparin-binding sites were observed in type IV collagen by the use of rotary shadowing: in the NC1 domain and at distances 100 and 300 nm from the NC1 domain. Scatchard analysis indicated different affinities for these sites. Electron microscopic analysis of heparin-type IV collagen interaction with increasing salt concentrations showed the different affinities to be NC1 greater than 100 nm greater than 300 nm. The NC1 domain bound specifically to chondroitin/dermatan sulfate side chains as well. This binding was observed at the electron microscope and in solid-phase binding assays (where chondroitin sulfate could compete for the binding of [3H]heparin to NC1-coated substrata). The triple helix-rich, rod-like domain of type IV collagen did not bind to chondroitin/dermatan sulfate side chains. In solid-phase binding assays only heparin could compete for the binding of [3H]heparin to this domain. In order to more precisely map potential heparin-binding sites in type IV collagen, we chemically synthesized 17 arginine- and lysine-containing peptides from the alpha 1(IV) and alpha 2(IV) chains. Three peptides from the known sequence of the alpha 1(IV) and alpha 2(IV) chains were shown to specifically bind heparin: peptide Hep-I (TAGSCLRKFSTM), from the alpha 1(NC1) chain, peptide Hep-II (LAGSCLARFSTM), a peptide corresponding to the same sequence in peptide Hep-I from the alpha 2 (NC1) chain, and peptide Hep-III (GEFYFDLRLKGDK) which contained an interruption of the triple helical sequence of the alpha 1(IV) chain at about 300 nm from the NC1 domain, were demonstrated to bind heparin in solid-phase binding assays and compete for the binding of [3H]heparin to type IV collagen-coated substrata. Therefore, each of these peptides may represent a potential heparin-binding site in type IV collagen. The mapping of the binding of heparin or related structures, such as heparan sulfate proteoglycan, to specific sequences of type IV collagen could help the understanding of several structural and functional properties of this basement membrane protein as well as interactions with other basement membrane and/or cell surface-associated macromolecules.  相似文献   

14.
The distances between the inositol 1,4,5-trisphosphate (IP(3))-binding sites of tetrameric IP(3) receptors were probed using dimers of IP(3) linked by poly(ethylene glycol) (PEG) molecules of differing lengths (1-8 nm). Each of the dimers potently stimulated (45)Ca(2+) release from permeabilized cells expressing predominantly type 1 (SH-SY5Y cells) or type 2 (hepatocytes) IP(3) receptors. The shortest dimers, with PEG linkers of an effective length of 1.5 nm or less, were the most potent, being 3-4-fold more potent than IP(3). In radioligand binding experiments using cerebellar membranes, the shortest dimers bound with highest affinity, although the longest dimer (8 nm) also bound with almost 4-fold greater affinity than IP(3). The affinity of monomeric IP(3) with only the PEG attached was 2-fold weaker than IP(3), confirming that the increased affinity of the dimers requires the presence of both IP(3) motifs. The increased affinity of the long dimer probably results from the linked IP(3) molecules binding to sites on different receptors, because the dimer bound with greater affinity than IP(3) to cerebellar membranes, where receptors are densely packed, but with the same affinity as IP(3) to purified receptors. IP(3) and the IP(3) dimers, irrespective of their length, bound with similar affinity to a monomeric IP(3)-binding domain of the type 1 IP(3) receptor expressed in bacteria. Short dimers therefore bind with increased affinity only when the receptor is tetrameric. We conclude that the four IP(3)-binding sites of an IP(3) receptor may be separated by as little as 1.5 nm and are therefore likely to be placed centrally in this large (25 x 25 nm) structure, consistent with previous work indicating a close association between the central pore and the IP(3)-binding sites of the IP(3) receptor.  相似文献   

15.
16.
Homogeneous noncompetitive assay of a protein in biological samples based on Förster-resonance-energy-transfer (FRET) was proposed by using its tryptophan residues as intrinsic donors and its specific fluorescent ligand as the FRET acceptor that was defined as an analytical FRET probe. Conjugate of a suitable fluorophore, which should have an excitation peak around 340 nm but an excitation valley around 280 nm, with a moiety binding to a protein of interest gave an analytical FRET probe to the protein. To test this method, N-biotinyl-N′-(1-naphthyl)-ethylenediamine (BNEDA) was used as an analytical FRET probe for homogeneous noncompetitive assay of streptavidin (SAV). The occurrence of FRET between the bound BNEDA and tryptophan residues was supported by the modeled geometry of the complex. By excitation at 280 nm, free BNEDA produced negligible fluorescence at 430 nm, but the bound BNEDA produced much higher stable fluorescence at 430 nm after 2 min of binding reaction. The competitive binding between BNEDA and biotin gave the dissociation constant of (16 ± 3) fM for BNEDA (n = 3). By excitation at 280 nm, fluorescence at 430 nm of reaction mixtures containing 32.0 nM BNEDA responded linearly to SAV subunit concentrations ranging from 0.40 to 30.0 nM with the desirable resistance to common interferences in biological samples. Therefore, by using tryptophan residue(s) in a protein of interest as intrinsic donor(s) and its fluorescent ligand as the corresponding FRET acceptor, this homogeneous noncompetitive assay of the protein in biological samples was effective and advantageous.  相似文献   

17.
Staphylococcus aureus pI258 CadC is an extrachromosomally encoded metalloregulatory repressor protein from the ArsR superfamily which negatively regulates the expression of the cad operon in a metal-dependent fashion. The metalloregulatory hypothesis holds that direct binding of thiophilic divalent cations including Cd(II), Pb(II), and Zn(II) by CadC allosterically regulates the DNA binding activity of CadC to the cad operator/promoter (O/P). This report presents a detailed characterization of the metal binding and DNA binding properties of wild-type CadC. The results of analytical ultracentrifugation experiments suggest that both apo- and Cd(1)-CadC are stable or weakly dissociable homodimers characterized by a K(dimer) = 3.0 x 10(6) M(-1) (pH 7.0, 0.20 M NaCl, 25.0 degrees C) with little detectable effect of Cd(II) on the dimerization equilibrium. As determined by optical spectroscopy, the stoichiometry of Cd(II) and Pb(II) binding is approximately 0.7-0.8 mol/mol of wild-type CadC monomer. Chelator (EDTA) competition binding isotherms reveal that Cd(II) binds very tightly, with K(Cd) = 4.3 (+/-1.8) x 10(12) M(-1). The results of UV-Vis and X-ray absorption spectroscopy of the Cd(1) complex are consistent with a tetrathiolate (S(4)) complex formed by four cysteine ligands. The (113)Cd NMR spectrum reveals a single resonance of delta = 622 ppm, consistent with an S(3)(N,O) or unusual upfield-shifted S(4) complex. The Pb(II) complex reveals two prominent absorption bands at 350 nm (epsilon = 4000 M(-1) cm(-1)) and 250 nm (epsilon = 41 000 M(-1) cm(-1)), spectral properties consistent with three or four thiolate ligands to the Pb(II) ion. The change in the anisotropy of a fluorescein-labeled oligonucleotide containing the cad O/P upon binding CadC and analyzed using a dissociable CadC dimer binding model reveals that apo-CadC forms a high-affinity complex [K(a) = (1.1 +/- 0.3) x 10(9) M(-1); pH 7.0, 0.40 M NaCl, 25 degrees C], the affinity of which is reduced approximately 300-fold upon the binding of a single molar equivalent of Cd(II) or Pb(II). The implications of these findings on the mechanism of metalloregulation are discussed.  相似文献   

18.
The interaction between a novel aromatic thiolato derivative from the family of DNA-intercalating platinum complexes, phenylthiolato-(2,2',2"-terpyridine)platinum(II)-[PhS(ter py)Pt+], and nucleic acids was studied by using viscosity, equilibrium-dialysis and kinetic measurements. Viscosity measurements with sonicated DNA provide direct evidence for intercalation, and show that at binding ratios below 0.2 molecules per base-pair PhS(terpy)Pt+ causes an increase in contour length of 0.2 nm per bound molecule. However, helix extension diminishes at greater extents of binding, indicating the existence of additional, non-intercalated, externally bound forms of the ligand. The ability of PhS(terpy)Pt+ to aggregate in neutral aqueous buffers at a range of ionic strengths and temperatures was assessed by using optical-absorption methods. Scatchard plots for binding to calf thymus DNA at ionic strength 0.01 (corrected for dimerization) are curvilinear, concave upward, providing further evidence for two modes of binding. The association constant decreases at higher ionic strengths, in accord with the expectations of polyelectrolyte theory, although the number of cations released per bound unipositive ligand molecule is substantially greater than 1. Stopped-flow kinetic measurements confirm the complexity of the binding reaction by revealing multiple bound forms of the ligand whose kinetic processes are both fast and closely coupled. Thermal denaturation of DNA radically alters the shapes of binding isotherms and either has little effect on, or enhances, the affinity of potential binding sites, depending on experimental conditions. Scatchard plots for binding to natural DNA species with differing nucleotide composition show that the ligand has a requirement for a single G X C base-pair at the highest-affinity intercalation sites.  相似文献   

19.
The single polypeptide chain of conalbumin strongly binds two Fe(III) or two Cu(II) ions to yield intense absorption in the visible region similar to that shown by the related protein transferrin. Comparison of the metal-ion-binding sites in the two proteins is made by exploiting the sensitivity to ligand geometry of circular dichroism (CD). For the Fe(III) proteins strong similarities of the CD spectra outweigh marginal differences. For Cu(II) conalbumin an additional negative extremum near 506 nm appears between two positive ones at 634 and 410 nm suggesting greater subtraction of oppositely signed CD components leading to lesser magnitudes for the two positive peaks than are found in Cu(II)-transferrin. The two Fe(III)-binding sites within conalbumin are compared by noting the strong similarities of the CD and MCD of proteins with Fe(III) in one site and Ga(III) in the other site, and vice versa, with the protein containing Fe(III) in both sites. Due to features of the amino acid sequences of the single protein chains, the four strong metal ion binding sites in conalbumin and transferrin cannot be identical in all particulars, yet CD spectra of their metal ion complexes are closely similar. From a study of model phenolate complexes and the wavelength maxima of visible absorption in the Fe(III), Cu(II), and Co(III) proteins near 465, 440, and 405 nm, respectively, these strong absorption bands are identified as ligand to metal ion electron-transfer transitions. It is suggested that tyrosyl residues are the donors in the electron transfer transitions and that they lock in the metal ions after being keyed into position by binding of bicarbonate or other anions.  相似文献   

20.
Metals are thought to play a role in the structure of many viruses. The crystal structure of the T=3 icosahedral cowpea chlorotic mottle virus (CCMV) suggests the presence of 180 unique metal-binding sites in the assembled protein cage. Each of these sites is thought to involve the coordination of the metal by five amino acids contributed from two adjacent coat protein subunits. We have used fluorescence resonance energy transfer (FRET), from tryptophan residues proximal to the putative metal-binding sites, to probe Tb(III) binding to the virus. Binding of Tb(III) was investigated on the wild-type virus and a mutant where the RNA binding ability of the virus was removed. Tb(III) binding was observed both in the wild-type virus (Kd=19 M) and the mutant (Kd=17 M), as monitored by the increase in Tb(III) fluorescence (545 nm) and concomitant decrease in tryptophan fluorescence (342 nm). Competitive binding experiments showed Ca(II) to have about 100-fold less affinity for the binding sites (Kd=1.97 mM). This is the first direct evidence of metal binding to the putative metal-binding sites, originally suggested from the crystal structure of CCMV.  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