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 78 毫秒
1.
The binding of Gd3+ ion to apoovotransferrin (apoOTf) was monitored by means of UV difference spectra in 0.01M Hepes, pH 7.4 at 25 degrees C. Used 2-p-toluidinylnaphthalene-6-sulfonate (TNS) as fluorescence probe the conformational changes of protein were studied while gadolinium ions bound to apoOTf. The results show that Gd3+ binding produces peaks at 244 and 294 nm that is the characteristic of binding at the apoOTf specific metal-binding sites. At 244 nm the molar absorptivity of Gd-apoOTf complex is (1.99+/-0.17)x10(4)cm(-1)M(-1). The apparent binding constants for the complexes of Gd3+ with apoovotransferrin are logK(1)=7.61+/-0.14 and logK(2)=4.96+/-0.26. A very large conformational change of apoovotransferrin appears when Gd3+ is bound to the N-terminal binding site. When Gd3+ is bound to C-terminal binding site there is less conformational change.  相似文献   

2.
Centrin is a member of the EF-hand superfamily that plays critical role in the centrosome duplication and separation. In the present paper, we characterized properties of metal ions binding to Euplotes octocarinatus centrin (EoCen) by fluorescence spectra and circular dichroism (CD) spectra. Changes of fluorescence spectra and alpha-helix contents of EoCen proved that Tb(3+) and Ca(2+) induced great conformational changes of EoCen resulting in exposing hydrophobic surfaces. At pH 7.4, Ca(2+) (and Tb(3+)) bond with EoCen at the ratio of 4:1. Equilibrium experiment indicated that Ca(2+) and Tb(3+) exhibited different binding capabilities for C- and N-terminal domains of protein. C-terminal domain bond with Ca(2+) or Tb(3+) approximately 100-fold more strongly than N-terminal. Aromatic residue-sensitized Tb(3+) energy transfer suggested that site IV bond to Tb(3+) or Ca(2+) more strongly than site III. Based on fluorescence titration curves, we reckoned the conditional binding constants of EoCen site IV quantitatively to be K(IV)=(1.23+/-0.51)x10(8)M(-1) and K(IV)=(6.82+/-0.33)x10(5)M(-1) with Tb(3+) and Ca(2+), respectively. Metal ions bond to EoCen in the order of IV>III>II, I.  相似文献   

3.
铝(Ⅲ)与脱铁伴清蛋白结合的紫外差光谱研究   总被引:2,自引:0,他引:2  
在 pH7.4、 0.1mol· L~(-1)N-2-羟乙基哌嗪- N′-2-乙磺酸( Hepes)及室温条件下,使用紫外吸收差光谱进行了铝(Ⅲ)对脱铁伴清蛋白的滴定。结果表明铝(Ⅲ)与脱铁伴清蛋白结合后其紫外差光谱在 238nm和 291nm处出现吸收峰。在 238nm处铝(Ⅲ)-脱铁伴清蛋白配合物的摩尔吸光系数是 (1.52± 0.04)× 10~4cm~(-1)· mol~(-1)· L。铝(Ⅲ)可占据脱铁伴清蛋白的两个金属离子结合部位,条件稳定常数是 lgK_N=11.21± 0.12,lgKC=9.53± 0.24。 N-端单铁伴清蛋白的紫外差光谱滴定表明,铝 ?优先占据脱铁伴清蛋白的 N端结合部位。  相似文献   

4.
李英奇  任学英  段新娥 《化学学报》2010,68(10):1027-1031
用TNS疏水探针研究了脱铁伴清蛋白(apoOTf)和不同形式的铕伴清蛋白(EuN-OTf, Eu2-OTf)的表面疏水暴露程度, 依次为apoOTf>EuN-OTf>Eu2-OTf, 表明Eu3+与脱铁伴清蛋白N端结合引起蛋白构象变化大, 与C端结合引起蛋白构象变化小. 此外, 由盐酸胍对3种蛋白变性实验, 发现Eu3+与脱铁伴清蛋白的结合稳定了蛋白的结构, 稳定性依次为apoOTf<EuN-OTf<Eu2-OTf. 离子强度效应也充分表明3种蛋白内部疏水基团相互作用依apoOTf<EuN-OTf<Eu2-OTf顺序增大, 稳定性也依次增大. 本研究对进一步探讨Eu3+的生物效应提供理论依据.  相似文献   

5.
In 0.01 M 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (Hepes), pH 7.4 and room temperature, the binding of neodymium to N,N'-ethylenebis[2-(o-hydroxyphenolic)glycine] (EHPG), or N,N'-di(2-hydroxybenzyl)ethylenediamine-N,N' diacetic acid (HBED) had been studied from 210 to 330 nm by means of difference UV spectra. Two peaks at 240 and 292 nm appear in difference UV spectra after neodymium binding to EHPG or HBED. The 1:1 stable complex can be confirmed from spectral titration curves. The molar extinction coefficient of Nd-EHPG and Nd-HBED complexes are Deltaepsilon(Nd-EHPG)=(12.93+/-0.21) x 10(3)cm(-1)M(-1), Deltaepsilon(Nd-HBED)=(14.45+/-0.51) x 10(5)cm(-1)M(-1) at 240 nm, respectively. Using EDTA as a competitor, the conditional equilibrium constants of the complexes are logK(Nd-EHPG)=11.89+/-0.09 and logK(Nd-HBED)=12.19+/-0.15, respectively. At the same conditions, fluorescence measurements show that neodymium binding to EHPG leads to a quenching of the fluorescence of EHPG at near 310 nm. However, there is no obvious fluorescence change of HBED at 318 nm with the binding of neodymium to HBED.  相似文献   

6.
Kinetics of metal ion exchange between citric acid and serum transferrin   总被引:1,自引:0,他引:1  
Harris WR  Wang Z  Brook C  Yang B  Islam A 《Inorganic chemistry》2003,42(19):5880-5889
The exchange of Fe(3+), Tb(3+), In(3+), Ga(3+), and Al(3+) between the C-terminal metal-binding site of the serum iron transport protein transferrin and the low-molecular-mass serum chelating agent citrate has been studied at pH 7.4 and 25 degrees C. The removal of Ga(3+), In(3+), and Al(3+) follows simple saturation kinetics with respect to the citrate concentration. In contrast, removal of both Fe(3+) and Tb(3+) shows a combination of saturation and first-order kinetic behavior with respect to the citrate concentration. The saturation component is consistent with a mechanism for metal release in which access to the bound metal is controlled by a rate-limiting conformational change in the protein. The first-order kinetic pathway is very rapid for Tb(3+), and this is attributed to a direct attack of the citrate on the Tb(3+) ion within the closed protein conformation. It is suggested that this pathway is more readily available for Tb(3+) because of the larger coordination number for this cation and the presence of an aquated coordination site in the Tb(3+)-CO(3)-Tf ternary complex. There is relatively little variation in the k(max) values for the saturation pathway for Tb(3+), Ga(3+), Al(3+), and In(3+), but the k(max) value for Fe(3+) is significantly smaller. It is suggested that protein interactions across the interdomain cleft of transferrin largely control the release of the first group of metal ions, while the breaking of stronger metal-protein bonds slows the rate of iron release. The rates of metal binding to apotransferrin are clearly controlled in large part by the hydrolytic tendencies of the free metal ions. For the more amphoteric metal ions Al(3+) and Ga(3+), there is rapid protein binding, and the addition of citrate actually retards this reaction. In contrast, the nonamphoteric In(3+) ion binds very slowly in the absence of citrate, presumably due to the rapid formation of polymeric In-hydroxo complexes upon addition of the unchelated metal ion to the pH 7.4 protein solution. The addition of citrate to the reaction accelerates the binding of In(3+) to apoTf, presumably by forming soluble, mononuclear In-citrate complexes.  相似文献   

7.
Solvent extraction of trivalent group 13 metal cations such as aluminum, gallium and indium with tripod quadridentate phenolic ligand, tris(2-hydroxy-3,5-dimethylbenzyl)amine (H(3)tdmba), was investigated as fundamental study for their mutual separation. Gallium was extracted almost quantitatively as Ga(tdmba) (logK(ex)=-6.66+/-0.06 on using chloroform as extraction solvent), whereas aluminum and indium were hardly extracted due to steric hindrance on complexation of them with the ligand. The extracted Ga species was estimated as trigonal bipyramidal complex with one H(2)O molecule. Furthermore, extractability of Ga was increased by changing the ligand to more acidic tris(5-chloro-2-hydroxy-3-methylbenzyl)amine (H(3)tcmba) (logK(ex)=-6.18+/-0.18 on using dichloroethane as extraction solvent).  相似文献   

8.
Binding Constants for Terbium(Ⅲ) with Chicken Apoovotransferrin   总被引:1,自引:0,他引:1  
IntroductionThetransferrinsareasuperfamilyofmetal bindingproteins ,whicharesingle chains ,80kDaglycoproteins ,characterizedbytheirabilitytobindirontightly(aneffectivebindingcon stantisabout 1 0 2 0 .7) ,butreversibly[1,2 ] .Serumtransferrin ,foundinbloodandextracellul…  相似文献   

9.
The interactions of yttrium with N,N'-ethylenebis[2-(o-hydroxyphenolic)glycine] (EHPG) and N,N'-di(2-hydroxybenzyl)ethylenediamine-N,N'-diacetic acid (HBED) are investigated by using UV difference and fluorescence spectra methods in 0.1M N-2-hydroxyethylpiperazine-N-2-ethanesulfonic acid (Hepes) at pH 7.4. Yttrium binding produces two UV difference peaks near 240 and 294 nm, respectively, that both are the characteristic of phenolic groups binding to yttrium. The molar extinction coefficient of Y-EHPG and Y-HBED are (15.7 +/- 0.40) x 10(3), (15.8 +/- 0.80) x 10(3)cm(-1)M(-1) at 240 nm, respectively. Using EDTA as a competitor the obtained conditional equilibrium constants of the complexes are logK(Y-EHPG) = 15.07 +/- 0.32 and logK(Y-HBED) = 15.18 +/- 0.26, respectively. However, the effects of yttrium binding on the fluorescence intensity of EHPG and HBED are quite different, the former showing a decrease but the latter an increase.  相似文献   

10.
Cyclic voltammetry (CV) and ultraviolet (UV) spectroscopy were used, for the first time, to study the interaction between aluminium(III) and calf thymus DNA under neutral pH conditions. Thus obtained data confirmed the existence of a relatively strong interaction between Al(III) and DNA. The binding site for aluminium(III) on DNA chains is not the bases, but the phosphate groups on the DNA backbones, the same as that for [Co(phen)3](3+/2+) that binds non-specifically and electrostatically to the deoxyribose phosphate backbone of DNA. When coexisting, Al(III) binds more favorably to DNA than [Co(phen)3](3+/2+), which implies the relatively strong binding of Al(III) to the phosphate backbone of DNA under neutral pH conditions. The nature of the binding of Al(III) to DNA is also discussed.  相似文献   

11.
The synthesis and characterization of two new macrocyclic ligands, the bis-macrocyclic compound 2,6-bis(1,4,13-triaza-7,10-dioxacyclopentadec-1-ylmethyl)phenol (L) and 38-methoxy-1,4,13,16,19,28-hexaaza-7,10,22,25-tetraoxatricyclo[14.14.7.1(32,36)]octatriconta-32,34,Delta(36,38)-triene (L1) are reported. Equilibrium studies of basicity and coordination properties toward metal ions such as Cu(II), Zn(II), Cd(II) and Pb(II) were performed for ligand by potentiometric measurements in aqueous solution (298.1 +/- 0.1 K, I= 0.15 mol dm(-3)). L behaves as a hexaprotic base (logK(1)= 10.93, logK(2)= 9.70, logK(3)= 8.79, logK(4)= 8.05, logK(5)= 6.83, logK(6)= 2.55). All metal ions form stable mono- and dinuclear complexes: logK(MLH(-1))= 25.61 for Cu(II), 15.37 for Zn(II), 12.58 for Cd(II) and 13.79 for Pb(II); logK(M(2)LH(-1))= 31.61 for Cu(II), 23.38 for Zn(II), 24.49 for Cd(II) and 23.68 for Pb(II). All these dinuclear species show a great tendency to add the OH(-) group: the equilibrium constant for the addition reaction was found to be logK(M(2)LH(-1)OH)= 4.77 for Cu(II), 5.66 for Zn(II), 2.8 for Cd(II) and 3.18 for Pb(II). In the case of Ni(II), kinetic inertness prevents the possibility of solution studies. The dinuclear solid adducts [Ni(2)H(-1)L(N(3))(3)].EtOH and [Cu(2)H(-1)L(N(3))](ClO(4))(2) were characterized by X-ray analysis.  相似文献   

12.
Thermodynamics of the binding of Ni(2+), Cu(2+) and Zn(2+) to bacitracin A(1) was studied by capillary electrophoresis measuring the peptide effective mobility at different pH in the presence of increasing concentration of the three ligands. The affinity follows the order Ni(2+) > Cu(2+) > Zn(2+), with association constant values of (2.3 +/- 0.1)x10(4), (4.9 +/- 0.2)x10(3), and (1.5 +/- 0.1)x10(3) M(-1), respectively. The only model able to rationalize mobility data implies that metal ion binds to the P(0) peptide form. Moreover, mobility values indicated a change of bacitracin A(1) acidic properties on Ni(2+) and Cu(2+) binding, with a shift of the pK(a) of N-terminal Ile-1 from 7.6 to about 5 and of the pK(a) of the delta-amino group of D-Orn-7 from 9.7 to about 7. Even though on Zn(2+) binding a shift of the N-terminal Ile-1 pK(a) was observed, restrictions in the pH range suitable for investigation, due to precipitation phenomena, did not allow establish if the shift of D-Orn-7 lateral chain pK(a) also occurred. Nonetheless, if present, the shift should be limited to the 7.8-9.7 range. Mobility data indicated that the Stokes radius of the complexes is ca. 3 A lower than that of the free peptide. The present results indicate that metal-ion binding to bacitracin A(1) is more complex than previously assumed.  相似文献   

13.
The ESR study of Cu(2+) doped calcium malonate dihydrate has been done at room temperature. Four magnetically in-equivalent sites for Cu(2+) have been observed. The spin-Hamiltonian parameters evaluated with the fitting of spectra to rhombic symmetry crystalline field are for Cu(2+) site (I): g(x)=2.0963+/-0.0002, g(y)=2.1316+/-0.0002, g(z)=2.4137+/-0.0002, A(x)=(32+/-2)x10(-4)cm(-1), A(y)=(34+/-2)x10(-4)cm(-1), A(z)=(49+/-2)x10(-4)cm(-1), for site (II): g(x)=2.0668+/-0.0002, g(y)=2.0800+/-0.0002, g(z)=2.3561+/-0.0002, A(x)=(34+/-2)x10(-4)cm(-1), A(y)=(36+/-2)x10(-4)cm(-1), A(z)=(51+/-2)x10(-4)cm(-1), for site (III): g(x)=2.0438+/-0.0002, g(y)=2.0623+/-0.0002, g(z)=2.2821+/-0.0002, A(x)=(34+/-2)x10(-4)cm(-1), A(y)=(36+/-2)x10(-4)cm(-1), A(z)=(53+/-2)x10(-4)cm(-1), and for site (IV): g(x)=2.0063+/-0.0002, g(y)=2.0241+/-0.0002, g(z)=2.2357+/-0.0002, A(x)=(35+/-2)x10(-4)cm(-1), A(y)=(37+/-2)x10(-4)cm(-1), A(z)=(54+/-2)x10(-4)cm(-1). The ground state wave function of Cu(2+) has also been determined. The g-anisotropy has been estimated and compared with the experimental value. Further with the help of optical study the nature of bonding of metal ion with different ligands in the complex has been discussed.  相似文献   

14.
The RNA helicase DbpA promotes RNA remodeling coupled to ATP hydrolysis. It is unique because of its specificity to hairpin 92 of 23S rRNA (HP92). Although DbpA kinetic pathways leading to ATP hydrolysis and RNA unwinding have been recently elucidated, the molecular (atomic) basis for the coupling of ATP hydrolysis to RNA remodeling remains unclear. This is, in part, due to the lack of detailed structural information on the ATPase site in the presence and absence of RNA in solution. We used high-field pulse ENDOR (electron-nuclear double resonance) spectroscopy to detect and analyze fine conformational changes in the protein's ATPase site in solution. Specifically, we substituted the essential Mg(2+) cofactor in the ATPase active site for paramagnetic Mn(2+) and determined its close environment with different nucleotides (ADP, ATP, and the ATP analogues ATPγS and AMPPnP) in complex with single- and double-stranded RNA. We monitored the Mn(2+) interactions with the nucleotide phosphates through the (31)P hyperfine couplings and the coordination by protein residues through (13)C hyperfine coupling from (13)C-enriched DbpA. We observed that the nucleotide binding site of DbpA adopts different conformational states upon binding of different nucleotides. The ENDOR spectra revealed a clear distinction between hydrolyzable and nonhydrolyzable nucleotides prior to RNA binding. Furthermore, both the (13)C and the (31)P ENDOR spectra were found to be highly sensitive to changes in the local environment of the Mn(2+) ion induced by the hydrolysis. More specifically, ATPγS was efficiently hydrolyzed upon binding of RNA, similar to ATP. Importantly, the Mn(2+) cofactor remains bound to a single protein side chain and to one or two nucleotide phosphates in all complexes, whereas the remaining metal coordination positions are occupied by water. The conformational changes in the protein's ATPase active site associated with the different DbpA states occur in remote coordination shells of the Mn(2+) ion. Finally, a competitive Mn(2+) binding site was found for single-stranded RNA construct.  相似文献   

15.
A helical peptide designed to present an all-leucine core upon folding has been shown to exhibit concentration-dependent helicity and to exist as an ill-defined equilibrium population of oligomers. In marked contrast, an identical peptide covalently modified with a 2,2'-bipyridyl group at the N terminus forms a stable three-stranded parallel coiled coil in the presence of transition metal ions. We have employed paramagnetic Ni(2+) and Co(2+) ions to stabilize the trimeric assembly and to exploit their shift and relaxation properties in NMR structural studies. We find that metal-ion binding and helix-bundle folding are tightly coupled. Surprisingly, the three-helix bundle exhibits a dynamic N-terminal region, and a well-structured C-terminal half. The spectra indicate the presence of a dual conformation for the bundle extending from the N terminus to residue 12. The structure of the two isomeric forms has been ascertained from interpretation of NOEs in the Ni(II) complex and (1)H pseudocontact shifts in the Co(II) complex. Two different facial isomers with distinct susceptibility tensors were identified. The bulky leucine side chain at position 3 in the peptide chain appears to play a role in the conformational variation at the N terminus.  相似文献   

16.
设计合成了蜂毒肽片断及其类似物: Mel15, Mel15(8F)和Mel15(7P), 这些多肽与钙调素有很强的结合力, 而且链段很短, 因此它们可作为钙调素可结合蛋白质的结合部位的模型。本文采用光谱法研究了它们与钙调素的相互作用。荧光发射光谱法结果表明, 多肽Mel15在与钙调素相互作用时, 肽链中的Trp基团的微环境变得更加疏水, 说明Mel15中的Trp残基可能与钙调素的疏水性表面靠近。紫外差谱测试表明, 只有当钙调素分子结合2个Ca^2^+后, 才可以与多肽Mel15(8F)结合。圆二色谱法研究表明, 多肽与钙调素结合后多肽分子和钙调素分子的α-螺旋结构的含量都被诱导而增加, 结合力越大, 则越多的残基被诱导形成α-螺旋结构。  相似文献   

17.

Background  

Escherichia coli DNA topoisomerase I binds three Zn(II) with three tetracysteine motifs which, together with the 14 kDa C-terminal region, form a 30 kDa DNA binding domain (ZD domain). The 67 kDa N-terminal domain (Top67) has the active site tyrosine for DNA cleavage but cannot relax negatively supercoiled DNA. We analyzed the role of the ZD domain in the enzyme mechanism.  相似文献   

18.
Circular dichroism and Fourier-transform infrared spectroscopies were used to compare the conformational mobility of 13-mer peptides covering the 317-329 region of the envelope protein hemagglutinin of human influenza A virus subtypes H1, H2 and H3 with that of their truncated deca- and nonapeptide analogs. These peptides were demonstrated to bind to the murine I-Ed major histocompatibility complex encoded class II and human HLA-B*2705 class I molecules. Despite the amino acid substitutions in the three 13-mer subtype sequences, no significant differences in the conformational properties could be shown. Deletion of the N-terminal three residues resulted in a shift to an increased alpha-helical conformer population in the 317-329 H1 peptide and the breakage of the 3(10) or weakly H-bonded (nascent) alpha-helix in the H2 and H3 peptides. The conformational change observed upon deletion did not influence the efficiency of I-Ed peptide interaction, however, the C-terminal Arg had a beneficial effect both on MHC class II and class I binding without causing any remarkable change in solution conformation.  相似文献   

19.
温度对铽(III)-转铁蛋白溶液构象的影响   总被引:2,自引:0,他引:2  
在pH 7.4,0.01 mol/L N-2-羟乙基哌嗪-N’-2-乙磺酸(Hepes)条件下,铽 (III)与N,N’-二(2-羟苄基)乙二胺-N,N’-二乙酸(HBED)结合并发生交换 相互作用使铽(III)荧光增强10~4倍,通过监测铽(III)545 nm荧光强度的变化 测定了Tb-HBED配合物的条件稳定常数是lgK = 14.30 ± 0.49;Tb-HBED配合物中 配体、铽(III)荧光强度均随着温度的升高而降低。在pH 7.4,0.01 mol/L Hepes条件下,Tb_N-apoTf-Tb_C配合物中蛋白质的荧光强度随着温度的升高而降 低,而能量受体铽(III)的荧光强度随着温度的升高而增强,主要源于铽(III) 与螺旋5色氨酸残基间的无辐射能量转移;当温度由0 ℃上升到55 ℃时,平均能量 转移效率AE值增加了29%,给体、受体间距离R有约4.2%的减小,温度变化引起 Tb_N-apoTf-Tb_C配合物大的构象变化;铽(III)与人血清脱铁转铁蛋白的结合使 蛋白质的变性温度降低。同样条件下,Tb_N-apoOTf-Tb_C配合物与Tb_N-apoTf- Tb_C配合物有所不同,虽然能量给体的荧光强度随着温度的增加而减小,但铽( III)荧光强度没有明显的增强;铽(III)对蛋白质的变性温度几乎没有影响。  相似文献   

20.
A series of dinickel(II) complexes with the 24-membered macrocyclic hexaazadithiophenol ligand H(2)L(Me) was prepared and examined. The doubly deprotonated form (L(Me))(2-) forms complexes of the type [(L(Me))Ni2II(mu-L')](n+) with a bioctahedral N(3)Ni(II)(mu-SR)(2)(mu-L')Ni(II)N(3) core and an overall calixarene-like structure. The bridging coordination site L' is accessible for a wide range of exogenous coligands. In this study L'=NO(3)(-), NO(2)(-), N(3)(-), N(2)H(4), pyrazolate (pz), pyridazine (pydz), phthalazine (phtz), and benzoate (OBz). Crystallographic studies reveal that each substrate binds in a distinct fashion to the [(L(Me))Ni(2)](2+) portion: NO(2)(-), N(2)H(4), pz, pydz, and phtz form mu(1,2)-bridges, whereas NO(3)(-), N(3)(-), and OBz(-) are mu(1,3)-bridging. These distinctive binding motifs and the fact that some of the coligands adopt unusual conformations is discussed in terms of complementary host-guest interactions and the size and form of the binding pocket of the [(L(Me))Ni(2)](2+) fragment. UV/Vis and electrochemical studies reveal that the solid-state structures are retained in the solution state. The relative stabilities of the complexes indicate that the [(L(Me))Ni(2)](2+) fragment binds anionic coligands preferentially over neutral ones and strong-field ligands over weak-field ligands. Secondary van der Waals interactions also contribute to the stability of the complexes. Intramolecular ferromagnetic exchange interactions are present in the nitrito-, pyridazine-, and the benzoato-bridged complexes where J=+6.7, +3.5, and +5.8 cm(-1) (H=-2 JS(1)S(2), S(1)=S(2)=1) as indicated by magnetic susceptibility data taken from 300 to 2 K. In contrast, the azido bridge in [(L(Me))Ni(2)(mu(1,3)-N(3))](+) results in an antiferromagnetic exchange interaction J=-46.7 cm(-1). An explanation for this difference is qualitatively discussed in terms of bonding differences.  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