首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
P B O'Hara  S H Koenig 《Biochemistry》1986,25(6):1445-1450
A human serum transferrin complex was prepared in which Gd(III) was substituted for Fe(III) at the two metal-binding sites. Characteristic changes upon metal binding in both the UV absorption of ligated tyrosines and the solvent proton longitudinal magnetic relaxation rates demonstrated 2/1 metal stoichiometry and pH-dependent binding constants. Binding studies were complicated both by binding of Gd(III) to nonspecific sites on transferrin at pH less than or equal to 7 and by complexation of the Gd(III) by the requisite bicarbonate anion at pH greater than or equal to 6.0. A unique Gd(III) electron spin resonance spectrum, with a prominent signal at g = 4.96, was observed for the specific Gd(III)-transferrin complex. The major features of this spectrum were fit successfully by a model Hamiltonian which utilized crystal field parameters similar to those determined for Fe(III) in transferrin [Aasa, R. (1970) J. Chem. Phys. 52, 3919-3924]. The magnetic field dependence of the solvent proton relaxation rate was measured as a function of both pH and metal ion concentration. An observed biphasic dependence of the relaxation rate on metal concentration is attributed to either sequential metal binding to the two iron-binding sites with different relaxation properties or random binding to two sites that are similar but show conformationally induced changes in relaxation properties as the second metal is bound. The increase in the solvent proton relaxation rate with pH is consistent with a model in which a proton of a second coordination sphere water molecule is hydrogen bonded to a metal ligand which becomes deprotonated at pH 8.5.  相似文献   

2.
The boronic functionalities on the outer surface of the Gd(III) bis(m-boroxyphenylamide)DTPA complex (Gd(III)L) enable it to bind to fructosamine residues of oxygenated glycated human adult hemoglobin. The formation of the macromolecular adduct can be assessed by NMR spectroscopy via observation of the enhancement of the solvent water proton relaxation rate. Unexpectedly, a strong binding interaction was also observed for the oxygenated unglycated human adult hemoglobin, eventually displaying a much higher relaxation enhancement. From relaxation rate measurements it was found that two Gd(III)L complexes interact with one hemoglobin tetramer (KD = 1.0 x 10(-5) M and 4.6 x 10(-4) M, respectively), whereas no interaction has been observed with monomeric hemoproteins. A markedly higher affinity of the Gd(III)L complex has been observed for oxygenated and aquo-met human adult hemoglobin derivatives with respect to the corresponding deoxy derivative. Upon binding, a net change in the quaternary structure of hemoglobin has been assessed by monitoring the changes in the high-resolution 1H-NMR spectrum of the protein as well as in the Soret absorption band. On the basis of these observations and the 11B NMR results obtained with the diamagnetic La(III)L complex, we suggest that the interaction between the lanthanide complex and deoxygenated, oxygenated, and aquo-met derivatives of human adult hemoglobin takes place at the 2, 3-diphosphoglycerate (DPG) binding site, through the formation of N-->B coordinative bonds at His143beta and His2beta residues of different beta-chains. The stronger binding to the oxygenated form is then responsible for a shift of the allosteric equilibrium toward the high-affinity R-state. Accordingly, Gd(III)L affinity for oxygenated human fetal hemoglobin (lacking His143beta) is significantly lower than that observed for the unglycated human adult tetramer.  相似文献   

3.
O Zak  P Aisen 《Biochemistry》1988,27(3):1075-1080
A wide variety of thermodynamic, kinetic, and spectroscopic studies have demonstrated differences between the two metal-binding sites of transferrin. In the present investigation, we have further assessed these differences with respect to the binding of gadolinium, evaluated by UV difference spectrophotometry, electron paramagnetic resonance (EPR) titration, EPR difference spectroscopy in conjunction with urea gel electrophoresis, and equilibrium dialysis. Combinations of these studies establish that only one site of the protein binds Gd(III) sufficiently firmly to be characterized. In order to reveal which of the two sites accepts Gd(III), we made use of monoferric transferrins preferentially loaded with Fe(III) at either site in EPR spectroscopic studies. Because of the overlap of signals, difference spectroscopy was required to distinguish resonances arising from Fe(III) and Gd(III) specifically complexed to the protein. When iron is bound to the C-terminal site, leaving the N-terminal site free for binding of gadolinium, the difference spectrum shows no evidence of specific binding. However, when iron is bound to the N-terminal site, the difference spectrum shows a resonance line at g' = 4.1 indicative of specific binding, thus implicating the C-terminal site in the binding of Gd(III). The effective stability constant for the binding of Gd(III) to this site of transferrin at pH 7.4 and ambient pCO2 is 6.8 X 10(6) M-1. At physiological pCO2, the formation of nonbinding carbonato complexes of Gd(III) precludes a substantial role for transferrin in the transport of the lanthanide in vivo.  相似文献   

4.
The gadolinium(III) chelates Gd(dtpaH2), Gd(hfac)3, Gd(tta)3 and Gd(qu)3 with dtpa=1,1,4,7,7-diethylenetriaminepentaacetate, hfac=hexafluoroacetylacetonate, tta=thenoyltrifluoroacetonate and qu=8-quinolinolate (or oxinate) show a phosphorescence under ambient conditions. While the UV emission of Gd(dtpaH2) at λmax=312 nm comes from a metal-centered ff state, the bluish (λmax=462 nm), green (λmax=505 nm) and red (λmax=650 nm) luminescence of Gd(hfac)3, Gd(tta)3 and Gd(qu)3, respectively, originates from the lowest-energy intraligand triplets.  相似文献   

5.
New solid complex compounds of La(III), Ce(III), Pr(III), Nd(III), Sm(III), Eu(III) and Gd(III) ions with morin were synthesized. The molecular formula of the complexes is Ln(C15H9O7)3 · nH2O, where Ln is the cation of lanthanide and n = 6 for La(III), Sm(III), Gd(III) or n = 8 for Ce(III), Pr(III), Nd(III) and Eu(III). Thermogravimetric studies and the values of dehydration enthalpy indicate that water occurring in the compounds is not present in the inner coordination sphere of the complex. The structure of the complexes was determined on the basis of UV-visible, IR, MS, 1H NMR and 13C NMR analyses. It was found that in binding the lanthanide ions the following groups of morin take part: 3OH and 4CO in the case of complexes of La, Pr, Nd, Sm and Eu, or 5OH and 4CO in the case of complexes of Ce and Gd. The complexes are five- and six-membered chelate compounds.  相似文献   

6.
The clinical application of macromolecular Gd(III) complexes as MRI contrast agents is impeded by their slow excretion and potential toxicity due to the release of Gd(III) ions caused by the metabolism of the agents. A polymer Gd(III) chelate conjugate with a cleavable spacer has been designed to solve this problem. Poly(l-glutamic acid)-cystamine-[Gd(III)-DOTA] was prepared by the conjugation of DOTA to PGA (MW = 50,000) via cystamine, a cleavable disulfide spacer, followed by the complexation with GdCl(3). A Gd(III) DOTA chelate derivative was readily released from the polymer conjugate in the incubation with cysteine, an endogenous plasma thiol. The conjugate produced significant MRI blood pool contrast enhancement in nude mice bearing OVCAR-3 human ovarian carcinoma xenographs. Less significant contrast enhancement was observed for a small molecular contrast agent, Gd(DTPA-BMA). The pharmacokinetic MRI study showed that the Gd(III) chelate from the conjugate accumulated in the urinary bladder in a similar kinetic pattern to Gd(DTPA-BMA), suggesting that the chelate was released by the endogenous thiols and excreted through renal filtration. The preliminary results suggest that this novel design has a great potential to solve the safety problem of macromolecular MRI contrast agents.  相似文献   

7.
Understanding how proteins are approached by surrounding molecules is fundamental to increase our knowledge of life at atomic resolution. Here, the surface accessibility of a multifunctional small protein, the archaeal protein Sso7d from Sulfolobus solfataricus, has been investigated by using TEMPOL and Gd(III)(DTPA-BMA) as paramagnetic probes. The DNA binding domain of Sso7d appears very accessible both to TEMPOL and Gd(III)(DTPA-BMA). Differences in paramagnetic attenuation profiles of (1)H-(15)N HSQC protein backbone amide correlations, observed in the presence of the latter paramagnetic probes, are consistent with the hydrogen bond acceptor capability of the N-oxyl moiety of TEMPOL to surface exposed Sso7d amide groups. By using the gadolinium complex as a paramagnetic probe a better agreement between Sso7d structural features and attenuation profile is achieved. It is interesting to note that the protein P-loop region, in spite of the high surface exposure predicted by the available protein structures, is not approached by TEMPOL and only partially by Gd(III)(DTPA-BMA).  相似文献   

8.
The binding of gadolinium [Gd(III)] to a homogeneous rabbit anti-(type-III pneumococcal polysaccharide) IgG (immunoglobulin G) and its Fab (N-terminal half of heavy and light chain) and Fc (C-terminal half of heavy-chain dimer) fragments was demonstrated by measurements of solvent-water proton relaxation rates in the appropriate Gd(III) solutions. At pH 5.5 the binding of Gd(III) to the Fc fragment is much tighter (KD approx. 5 micronM) than binding to the Fab fragment (KD approx. 250 micronM). The binding of Gd(III) to the whole IgG molecule (KD approx. 4 micronM) is very similar to that for the Fc fragment alone. This specificity of binding to the Fc region allows the use of Gd(III) as a probe of the Fc conformation. The environment of the Gd(III) in the Fc region of whole IgG is not affected by the presence of octasaccharide derived by hydrolysis of type-III pneumococcal polysaccharide, but the corresponding 28-unit saccharide does cause detectable changes. The addition of 16-unit saccharide to anti-(SIII polysaccharide) IgG in the presence of Gd(III) does not change the solvent water proton relaxation rate, although aggregation does occur. The effects of the 28-unit saccharide may be explained therefore by a change in the tumbling time of the IgG. From a study of the effect of various antigen/antibody ratios, it is concluded that the 28-unit-saccharide-induced changes in the Gd(III) environment in the Fc region are caused by the specific geometrical structure of the antigen-antibody complexes formed, and not simply by occupancy of the combining sites on the antibody.  相似文献   

9.
Two novel Gd(III) complexes with functionalised polyaminocarboxylate macrocycles, 1,4,7-tris(carboxymethyl)-9,24-dioxo-14,19-dioxa-1,4,7,10,23- pentaazacyclododecane (L(1)) and 1,4,7-tris(carboxymethyl)-9,25-dioxo-14,17,20-trioxa-1,4,7,10,23- pentaazacyclotridecane (L(2)), were prepared in good yield. Their potential use as magnetic resonance imaging (MRI) contrast agents (CAs) was evaluated by investigating their relaxation behaviour as a function of pH, temperature and magnetic field strength. The 1/T(1) proton relaxivities at 20 MHz and 25 degrees C of GdL(1) (5.87 mM(-1) s(-1)) and GdL(2) (6.14 mM(-1) s(-1)) were found to be significantly higher than the clinically used Gd 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid (Gd(DOTA)(-)) and Gd diethylenetriaminepentaethanoic acid (Gd(DTPA)(2-)). The complexes possess one water molecule in the inner coordination sphere whose mean residence lifetime was estimated to be 1.1 and 1.5 micros at 25 degrees C by variable temperature (VT) (17)O NMR spectroscopy.  相似文献   

10.
1. The interactions of lanthanide metals and dinitrophenyl spin-label haptens with the Fv fragment of the mouse myeloma protein MOPC 315 were investigated by the techniques of fluorescence, e.s.r. (electron spin resonance) and high-resolution n.m.r. (nuclear magnetic resonance). 2. The protein fluorescence of Fv fragment at 340nm is quenched by the haptens (fluorescence enhancement, epsilon=0.15) and enhanced by Gd(III) (epsilon=1.14) and other lanthanides. The binding of the haptens studied here is insensitive to pH in the range 5.5-7.0 (dissociation constant KH=0.3-1.0 muM) and shows 1:1 stoicheiometry. The binding of Gd(III) also shows 1:1 stoicheiometry, but is pH-dependent; the binding constant (KM) varies from 10 muM at pH7.0 to 700 muM at pH4.8. La(III) binding is less sensitive to pH. The pH-dependences of the metal-binding constants imply that a group in the protein with pKa greater than or equal to 6.2 is involved in the binding, and probably also other groups with lower pKa values. 3. The apparent binding of the haptens is weakened about 20-fold by Gd(III), and vice versa. An equilibrium scheme involving a ternary complex with an interaction between the two binding sites is derived in Appendix I to explain the experimental results at two pH values. 4. Time-dependent fluorescence changes are observed in the presence of Gd(III) at pH5.5. A two-state kinetic scheme involving a 'slow' conformational change in the Fv fragment is derived in Appendix II to explain this time-dependence. This scheme is consistent with the antagonistic equilibrium behaviour. 5. The e.s.r. changes in the spin-label haptens on binding to Fv fragment and on the subsequent addition of lanthanides are consistent with the binding scheme for haptens and lanthanides proposed from the fluorescence studies. A difference between the limiting quenching of the e.s.r. signal from the bound haptens in the presence of saturating concentrations of Gd(III) and La(III) is attributed to dipolar interactions between bound Gd(III) and the nitroxide moiety of the bound hapten. The residual quenching with Gd(III) allows an estimate of 1.2nm to be made for the distance between the two paramagnetic centres. 6. The 270 MHz proton difference spectrum of the Fv fragment resulting from the addition of La(III) suggests that any metal-induced conformational changes are small and involve relatively few amino acid residues on the Fv fragment...  相似文献   

11.
Immobilized metal ion affinity chromatography has been used to demonstrate and partially characterize Fe(III) binding sites on apoferritin. Binding of Fe(III) to these sites is influenced by pH, but not affected by high ionic strength. These results suggest that both ionic and coordinate covalent interactions are important in the formation of the Fe(III): apoferritin complex. This is, to our knowledge, the first demonstration of direct Fe(III) binding to apoferritin. Other immobilized metal ions, including Zn(II), Ni(II), Cu(II), Cr(III), Co(II), and Tb(III), displayed little or no adsorption of apoferritin. The analytical technique of immobilized metal ion affinity chromatography also shows great promise in the purification of apoferritin, ferritin, and other iron-binding proteins.  相似文献   

12.
The interaction of lectin isolated from rice (Oryza sativa) embryos with N-acetylglucosaminides was studied by equilibrium dialysis and fluorescence. Equilibrium dialysis with 4-methylumbelliferyl-(GlcNac)2 showed that rice lectin (Mr 38000) contains four equivalent saccharide-binding sites. Addition of the N-acetylglucosaminides GlcNac, (GlcNac)2 and (GlcNac)3 enhanced the intrinsic fluorescence of rice lectin and this was accompanied by a 10nm blue-shift of its maximum fluorescence with (GlcNac)2 and (GlcNac)3. These changes in intensity allowed determination of the association constants, which increased with the number of saccharide units: at 20 degrees C, Ka = (1.3 +/- 0.1) X 10(3), (5.1 +/- 0.4) X 10(4) and (2.6 +/- 0.1) X 10(5) M-1 for GlcNac, (GlcNac)2 and (GlcNac)3 respectively. The binding enthalpy, delta H0, for the three glucosaminides were very low and ranged from -12.1 to -20.6 kJ X mol-1. The results are compared with those obtained with wheat-germ agglutinin, another GlcNac-specific gramineaous lectin.  相似文献   

13.
《Inorganica chimica acta》1988,154(2):221-224
Polynuclear sulfur bridged complexes where the neutral complex tris(2-aminoethanethiolato)cobalt(III) acts as a tridentate ligand to rhodium(III), iridium(III) and osmium(III) have been prepared. These complexes have been characterized by electronic spectroscopy, vibrational spectroscopy and nuclear magnetic resonance spectroscopy. Along with the previously prepared complexes of iron(III), ruthenium(III) and cobalt(III), these complexes form two series of complexes with the group 8 and group 9 elements from all three transition series.  相似文献   

14.
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   

15.
Intramolecular proton transfer of heme-copper oxidases is performed via the K- and the transmembrane D-channels. A carboxyl group conserved in a subgroup of heme-copper oxidases, located within the D-channel close to the binuclear center (=glutamic acid-286 in cytochrome bo(3) from Escherichia coli) is essential for proton pumping. Upon electron transfer to the fully oxidized (FO) enzyme, this amino acid has been shown to undergo a cyanide-independent environmental change. The redox-induced environmental transition of glutamic acid-286 is preserved in the site-directed mutant Y288F, which has lost its Cu(B) binding capacity. Furthermore, the mixed-valence (MV) redox state of cytochrome bo(3) (in which Cu(B) and high-spin heme are reduced, whereas the low-spin heme stays oxidized) was prepared by anaerobic exposure of the protein to carbon monoxide. This complex was converted (i) to the FO state by reaction with the caged dioxygen donor mu-peroxo) (mu-hydroxo) bis [bis (bipyridyl) cobalt (III)] and (ii) to the fully reduced (FR) state via caged electron donors; the environmental change of glutamic acid-286 could be observed only upon reduction. Taken together, these results from two different lines of evidence clearly show that the redox transition of the low-spin heme b center alone triggers the change in the chemical environment of this acidic side chain. It is suggested that glutamic acid-286 is a kinetic enhancer of proton translocation, which is energetically favoured in mesophilic oxidases.  相似文献   

16.
The substitution of trivalent lanthanide ions for Ca(II) in the Ca(II)-DEPENDENT ACTIVATION OF BOVINE Factor X by the coagulant protein of Russell's viper venom was studied at pH 6.8. Factor X contains two high affinity metal binding sites which bind Gd(III), Sm(III), and Yb(III) with a Kd of about 4 X 10-7 M and four to six lower affinity metal binding sites which bind Gd(III), Sm(III) with a Kd of about 1.5 X 10-5M. In comparison, 1 mol of Factor X binds 2 mol of Ca(II) with a Kd of 3 X 10-4M and weakly binds many additional Ca(II) ions. No binding of Gd(III) to the venom protein was observed. Dy(III), Yb(III), Tb(III), Gd(III), Eu(III), La(III), AND Nd(III) cannot substitute for Ca(II) in the Ca(II)-dependent activation of Factor X by the venom protein at pH 6.8. Kinetic data consistent with the models of competitive inhibition of Ca(II) by Nd(III) yielded a Ki of 1 to 4 X 10-6M. The substitution of lanthanide ions for Ca(II) to promote protein complex formation of Factor X-metal-venom protein without the activation of Factor X facilitated the purification of the coagulant protein from crude venom by affinity chromatography. Using a column containing Factor X covalently bound to agarose which was equilibrated in 10 mM Nd(III), Tb(III), Gd(III), or La(III), the coagulant protein was purified 10-fold in 40% yield from crude venom and migrated as a single band on gel electrophoresis in sodium dodecyl sulfate. These data suggest that lanthanide ions complete with Ca(II) for the metal binding sites of Factor X and facilitate the formation of a nonproductive ternary complex of venom protein-Factor X-metal. Tb(III) fluorescence, with emission maxima at 490 and 545 nm, is enhanced 10,000-fold in the presence of Factor X. The study of the participation of an energy donor intrinsic to Factor X in energy transfer to Tb(III) may be useful in the characterization of the metal binding sites of Factor X.  相似文献   

17.
The gadolinium species present in a rat kidney following intravenous administration of a gadolinium-based magnetic resonance contrast agent (Optimark?, Gadoversetamide injection) to a rat was examined in the present study. The major gadolinium species in the supernatant of the rat kidney tissue extracts was determined by reversed-phase liquid chromatography with online inductively coupled plasma optical emission spectrometry (HPLC-ICP-OES). The identity of the compound was established by liquid chromatography electrospray ionization mass spectrometry (LC-ESI-MS) detection. The principal gadolinium(III) complex in a rat kidney tissue extract was identified as Gd-DTPA-BMEA 24 Hrs and 7 days after a single intravenous injection of Optimark? (gadoversetamide; Gd-DTPA-BMEA) at a dose of 5 mmol Gd/kg body weight. The study demonstrated for the first time the feasibility of the use of two complementary techniques, HPLC-ICP-OES and HPLC-ESI-MS to study the in vivo behavior of gadolinium-based magnetic resonance contrast media.  相似文献   

18.
The speciation and distribution of Gd(III) in human interstitial fluid was studied by computer simulation. The results show that at the background concentration, all the Gd(III) species are soluble and no precipitates appear. However as the total concentration of Gd(III) rises above 2.610 × 10–9 mol/l,the insoluble species become predominant. GdPO4 is formed first as a precipitate and then Gd2(CO3)3. Among soluble species, free Gd(III), [Gd(HSA)], [Gd(Ox)] and the ternary complexes of Gd(III) with citrate as the primary ligand are main species when the total concentration of Gd(III) is below 2.074 × 10–2 mol/l. With the total concentration of Gd(III) further rising, [Gd3(OH)4] begins to appear and gradually becomes a predominant species.  相似文献   

19.
The cerium (III) glutathione complex was synthesized by the redox reaction of cerium (IV) with glutathione reduced (GSH) in aqueous solution. The Job‐plots indicate an ML (L = GSSG) stoichiometry of the complex. The fluorescent properties of the compound were investigated. The as‐prepared complex showed the characteristic maximum emission spectra of Ce(III) at 350 nm (λex = 255 nm). The fluorescence results show that the Ce(IV) ions are first reduced to Ce(III), and then form Ce(III) complex after reacting with GSH. The complex was characterized by element analysis and FT‐IR spectra; the stability of the complex was analyzed by cyclic voltammeters and DSC‐TG as well. Finally, Ce(IV) was successfully employed to determine the concentrations of GSH in the presence of GSSG, in which the fluorescence intensities are proportional to the concentrations of GSH in the range of 1–100 nM with the detection limit of 0.05 nM of GSH, without interference from the presence of GSSG. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
 A novel heptacoordinating ligand consisting of a thirteen-membered tetraazamacrocycle containing the pyridine ring and bearing three methylenephosphonate groups (PCTP-[13]) has been synthesized. Its Gd(III) complex displays a remarkably high longitudinal water proton relaxivity (7.7 mM–1 s–1 at 25  °C, 20 MHz and pH 7.5) which has been accounted for in terms of contributions arising from (1) one water molecule bound to the metal ion, (2) hydrogen-bonded water molecules in the second coordination sphere, or (3) water molecules diffusing near the paramagnetic chelate. Variable-temperature 17O-NMR transverse relaxation data indicate that the residence lifetime of the metal-bound water molecule is very short (8.0 ns at 25  °C) with respect to the Gd(III) complexes currently considered as contrast agents for magnetic resonance imaging. Furthermore, GdPCTP-[13] interacts with human serum albumin (HSA), likely through electrostatic forces. By comparing water proton relaxivity data for the GdPCTP-[13]-HSA adduct, measured as a function of temperature and magnetic field strength, with those for the analogous adduct with GdDOTP (a twelve-membered tetraaza macrocyclic tetramethylenephosphonate complex lacking a metal-bound water molecule), it has been possible to propose a general picture accounting for the main determinants of the relaxation enhancement observed when a paramagnetic Gd(III) complex is bound to HSA. Basically, the relaxation enhancement in these systems arises from (1) water molecules in the hydration shell of the macromolecule and protein exchangeable protons which lie close to the interaction site of the paramagnetic complex and (2) the metal bound water molecule(s). As far as the latter contribution is concerned, the interaction with the protein causes an elongation of the residence lifetime of the metal-bound water molecule, which limits, to some extent, the potential relaxivity enhancement expected upon the binding of the paramagnetic complex to HSA. Received: 27 January 1997 / Accepted: 12 May 1997  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