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.
In vitro channel activity of the C-terminal colicin E1 channel polypeptide under conditions of variable electrostatic interaction with synthetic lipid membranes showed distinct maxima with respect to pH and membrane surface potential. The membrane binding energy was determined from fluorescence quenching of the intrinsic tryptophans of the channel polypeptide by liposomes containing N-trinitrophenyl-phosphatidylethanolamine. Maximum in vitro colicin channel activity correlated with an intermediate magnitude of the electrostatic interaction. For conditions associated with maximum activity (40% anionic lipid, I = 0.12 M, pH 4.0), the free energy of binding was delta G approximately -9 kcal/mol, with nonelectrostatic and electrostatic components, delta Gnel approximately -5 kcal/mol and delta Gel approximately -4 kcal/mol, and an effective binding charge of +7 at pH 4.0. Binding of the channel polypeptide to negative membranes at pH 8 is minimal, whereas initial binding at pH 4 followed by a shift to pH 8 causes only 3-10% reversal of binding, implying that it is kinetically trapped, probably by a hydrophobic interaction. It was inferred that membrane binding and insertion involves an initial electrostatic interaction responsible for concentration and binding to the membrane surface. This is followed by insertion into the bilayer driven by hydrophobic forces, which are countered in the case of excessive electrostatic binding.  相似文献   

2.
Recently, a new nonpathogenic transthyretin (TTR) variant-TTR R104H (TTR H104)-has been described in heterozygotic and compound heterozygotic individuals from a Japanese family with familial amyloidotic polyneuropathy (FAP). The compound heterozygotic individual, a carrier of TTR V30M (TTR M30) and TTR R104H (TTR M30/H104) presented a very mild form of FAP with slow progression of the disease. TTR and retinol binding protein (RBP) levels were found to be increased in serum from TTR H104 carriers. These characteristics are very similar to those found in compound heterozygotic carriers of TTR V30M-T119M (TTR M30/M119). To structurally compare these variants, we performed stability and thyroxine (T(4)) binding studies. TTR M30/H104 showed an increased resistance to dissociation into monomers similar to TTR M30/M119. This suggests that the His104 substitution has the same stabilizing effect on tetrameric TTR as the Met119 substitution. Concerning T(4) binding, TTR H104 presents a T(4) binding affinity lower than that of TTR M119, but still higher than normal TTR. However, TTR from the compound heterozygotic carrier of TTR M30/H104 presented a T(4) binding affinity lower than normal. The results indicate that the His 104 substitution induces structural alterations that increase the stability of the tetramer in compound heterozygotes for TTR M30 despite a lower affinity for T(4) binding. Thus, stability of TTR and binding affinity for T(4) may not be related. More detailed characterization of these variants is needed to clarify the structural alterations responsible for their increased stability.  相似文献   

3.
Transthyretin (TTR) is an extracellular transport protein involved in the distribution of thyroid hormones and vitamin A. So far, TTR has only been found in vertebrates, of which piscine TTR displays the lowest sequence identity with human TTR (47%). Human and piscine TTR bind both thyroid hormones 3,5,3'-triiodo-l-thyronine (T(3)) and 3,5,3',5'-tetraiodo-l-thyronine (thyroxine, T(4)). Human TTR has higher affinity for T(4) than T(3), whereas the reverse holds for piscine TTR. X-ray structures of Sparus aurata (sea bream) TTR have been determined as the apo-protein at 1.75 A resolution and bound to ligands T(3) and T(4), both at 1.9 A resolution. The apo structure is similar to human TTR with structural changes only at beta-strand D. This strand forms an extended loop conformation similar to the one in chicken TTR. The piscine TTR.T(4) complex shows the T(4)-binding site to be similar but not identical to human TTR, whereas the TTR.T(3) complex shows the I3' halogen situated at the site normally occupied by the hydroxyl group of T(4). The significantly wider entrance of the hormone-binding channel in sea bream TTR, in combination with its narrower cavity, provides a structural explanation for the different binding affinities of human and piscine TTR to T(3) and T(4).  相似文献   

4.
During vertebrate evolution, the N-terminal region of transthyretin (TTR) subunit has undergone a change in both length and hydropathy. This was previously shown to change the binding affinity for thyroid hormones (THs). However, it was not known whether this change affects other functions of TTR. In the present study, the effect of these changes on the binding of TTR to retinol-binding protein (RBP) was determined. Two wild-type TTRs from human and Crocodylus porosus, and three chimeric TTRs, including a human chimeric TTR in which its N-terminal sequence was changed to that of C. porosus TTR (croc/huTTR) and two C. porosus chimeric TTRs (hu/crocTTR in which its N-terminal sequence was changed to that of human TTR and xeno/crocTTR in which its N-terminal sequence was changed to that of Xenopus laevis TTR), were analyzed for their binding to human RBP by native-PAGE followed by immunoblotting and a chemilluminescence assay. The K(d) of human TTR was 30.41 ± 2.03 μm, and was similar to that reported for the second binding site, whereas that of crocodile TTR was 2.19 ± 0.24 μm. The binding affinities increased in croc/huTTR (K(d) = 23.57 ± 3.54 μm) and xeno/crocTTR (K(d) = 0.61 ± 0.16 μm) in which their N-termini were longer and more hydrophobic, but decreased in hu/crocTTR (K(d) = 5.03 ± 0.68 μm) in which its N-terminal region was shorter and less hydrophobic. These results suggest an influence of the N-terminal primary structure of TTR on its function as a co-carrier for retinol with RBP.  相似文献   

5.
The binding of alpha-, omega-amino acids, which are important effectors of human plasminogen activation, to the isolated kringle 4 (K4) peptide region of this protein has been investigated by high sensitivity titration calorimetry. The titration curve of the heat changes accompanying binding of the widely employed ligand, epsilon-aminocaproic acid (EACA), to K4 were deconvoluted to yield the following binding characteristics: n = 0.87 +/- 0.08 mol/mol; Ka = 3.82 +/- 0.37 x 10(4) M-1; delta H = -4.50 +/- 0.22 kcal/mol; delta S = 6.01 +/- 0.7 entropy units; and delta G = 6.29 +/- 0.06 kcal/mol. Here, both delta H and delta S provide the driving force of the interaction, with both hydrogen bonds and hydrophobic interactions, the latter which may result from an induced conformational change in K4 upon ligand binding, as well as possible alterations in peptide-bound water structure, providing the stabilizing forces for complex formation. The thermodynamic binding parameters were not greatly influenced by pH between the values of 5.5 and 8.2, suggesting that titratable groups on K4 in this pH region did not influence the binding. Investigations of the binding properties of structural analogues of EACA to K4 demonstrated that definable steric requirements existed for a maximal interaction, with spacing between the functional groups on EACA, as well as a hydrophobic region of this molecule, being important. This rapid and reliable method for measuring all thermodynamic parameters of formation of this complex at a given temperature can now be employed to investigate this important interaction with a wide variety of kringles and modified kringles to provide a more complete understanding of the necessary factors for this binding to occur.  相似文献   

6.
Two colchicine analogs with modifications only in the C ring are better inhibitors than colchicine of cell growth and tubulin polymerization. Radiolabeled thiocolchicine (with a thiomethyl instead of a methoxy group at position C-10) and N-acetylcolchinol O-methyl ether (NCME) (with a methoxy-substituted benzenoid instead of the methoxy-substituted tropone C ring) were prepared for comparison with colchicine. Scatchard analysis indicated a single binding site with KD values of 1.0-2.3 microM. Thiocolchicine was bound 2-4 times as rapidly as colchicine, but the activation energies of the reactions were nearly identical (18 kcal/mol for colchicine, 20 kcal/mol for thiocolchicine). NCME bound to tubulin in a biphasic reaction. The faster phase was 60 times as fast as colchicine binding at 37 degrees C, and a substantial reaction occurred at 0 degrees C. The rate of the faster phase of NCME binding changed relatively little as a function of temperature, so the activation energy was only 7.0 kcal/mol. Dissociation reactions were also evaluated, and at 37 degrees C the half-lives of the tubulin-drug complexes were 11 min for NCME, 24 h for thiocolchicine, and 27 h for colchicine. Relative dissociation rates as a function of temperature varied little among the drug complexes. Activation energies for the dissociation reactions were 30 kcal/mol for thiocolchicine, 27 kcal/mol for NCME, and 24 kcal/mol for colchicine. Comparison of the activation energies of association and dissociation yielded free energies for the binding reactions of -20 kcal/mol for NCME, -10 kcal/mol for thiocolchicine, and -6 kcal/mol for colchicine. The greater effectiveness of NCME and thiocolchicine as compared with colchicine in biological assays probably derives from their more rapid binding to tubulin and the lower free energies of their binding reactions.  相似文献   

7.
Snider MJ  Wolfenden R 《Biochemistry》2001,40(38):11364-11371
Kinetic measurements have shown that substantial enthalpy changes accompany substrate binding by cytidine deaminase, increasing markedly as the reaction proceeds from the ground state (1/K(m), DeltaH = -13 kcal/mol) to the transition state (1/K(tx), DeltaH = -20 kcal/mol) [Snider, M. J., et al. (2000) Biochemistry 39, 9746-9753]. In the present work, we determined the thermodynamic changes associated with the equilibrium binding of inhibitors by cytidine deaminase by isothermal titration calorimetry and van't Hoff analysis of the temperature dependence of their inhibition constants. The results indicate that the binding of the transition state analogue 3,4-dihydrouridine DeltaH = -21 kcal/mol), like that of the transition state itself (DeltaH = -20 kcal/mol), is associated with a large favorable change in enthalpy. The significantly smaller enthalpy change that accompanies the binding of 3,4-dihydrozebularine (DeltaH = -10 kcal/mol), an analogue of 3,4-dihydrouridine in which a hydrogen atom replaces this inhibitor's 4-OH group, is consistent with the view that polar interactions with the substrate at the site of its chemical transformation play a critical role in reducing the enthalpy of activation for substrate hydrolysis. The entropic shortcomings of 3,4-dihydrouridine, in capturing all of the free energy involved in binding the actual transition state, may arise from its inability to displace a water molecule that occupies the binding site normally occupied by product ammonia.  相似文献   

8.
Deposition of amorphous aggregates and fibrils of transthyretin (TTR) in leptomeninges and subarachnoid vessels is a characteristic of leptomeningeal amyloidosis (LA), a currently untreatable cerebral angiopathy. Herein, we report the X-ray structure of the A25T homotetramer of TTR, a natural mutant described in a patient with LA. The structure of A25T-TTR is indistinguishable from that of wild-type TTR (wt-TTR), indicating that the difference in amyloidogenicity between A25T-TTR and wt-TTR cannot be ascribed to gross structural differences. Using pressure-induced dissociation of the tetramer, we show that A25T-TTR is 3 kcal/mol less stable than L55P-TTR, the most aggressive mutant of TTR described to date. After incubation for 15 days at 37 °C (pH 7.3), A25T-TTR forms mature amyloid fibrils. To mimic the environment in which TTR aggregates, we investigated aggregation in cerebrospinal fluid (CSF). Unlike L55P-TTR, A25T-TTR rapidly forms amyloid aggregates in CSF that incorporated several protein partners. Utilizing a proteomics methodology, we identified 19 proteins that copurified with A25T-TTR amyloid fibrils. We confirmed the presence of proteins previously identified to be associated with TTR aggregates in biopsies of TTR amyloidosis patients, such as clusterin, apolipoprotein E, and complement proteins. Moreover, we identified novel proteins, such as blood coagulation proteins. Overall, our results revealed the in vitro characterization of TTR aggregation in a biologically relevant environment, opening new avenues of investigation into the molecular mechanisms of LA.  相似文献   

9.
Here, we compare an antibody with the highest known engineered affinity (K(d)=270 fM) to its high affinity wild-type (K(d)=700 pM) through thermodynamic, kinetic, structural, and theoretical analyses. The 4M5.3 anti-fluorescein single chain antibody fragment (scFv) contains 14 mutations from the wild-type 4-4-20 scFv and has a 1800-fold increase in fluorescein-binding affinity. The dissociation rate is approximately 16,000 times slower in the mutant; however, this substantial improvement is offset somewhat by the association rate, which is ninefold slower in the mutant. Enthalpic contributions to binding were found by calorimetry to predominate in the differential binding free energy. The crystal structure of the 4M5.3 mutant complexed with antigen was solved to 1.5A resolution and compared with a previously solved structure of an antigen-bound 4-4-20 Fab fragment. Strikingly, the structural comparison shows little difference between the two scFv molecules (backbone RMSD of 0.6A), despite the large difference in affinity. Shape complementarity exhibits a small improvement between the variable light chain and variable heavy chain domains within the antibody, but no significant improvement in shape complementarity of the antibody with the antigen is observed in the mutant over the wild-type. Theoretical modeling calculations show electrostatic contributions to binding account for -1.2 kcal/mol to -3.5 kcal/mol of the binding free energy change, of which -1.1 kcal/mol is directly associated with the mutated residue side-chains. The electrostatic analysis reveals several mechanistic explanations for a portion of the improvement. Collectively, these data provide an example where very high binding affinity is achieved through the cumulative effect of many small structural alterations.  相似文献   

10.
N Noy  Z J Xu 《Biochemistry》1990,29(16):3888-3892
Retinol (vitamin A alcohol) is a hydrophobic compound and distributes in vivo mainly between binding proteins and cellular membranes. To better clarify the nature of the interactions of retinol with these phases which have a high affinity for it, the thermodynamic parameters of these interactions were studied. The temperature-dependence profiles of the binding of retinol to bovine retinol binding protein, bovine serum albumin, unilamellar vesicles of dioleoylphosphatidylcholine, and plasma membranes from rat liver were determined. It was found that binding of retinol to retinol binding protein is characterized by a large increase in entropy (T delta S degrees = +10.32 kcal/mol) and no change in enthalpy. Binding to albumin is driven by enthalpy (delta H degrees = -8.34 kcal/mol) and is accompanied by a decrease in entropy (T delta S degrees = -2.88 kcal/mol). Partitioning of retinal into unilamellar vesicles and into plasma membranes is stabilized both by enthalpic (delta H degrees was -3.3 and -5.5 kcal/mol, respectively) and by entropic (T delta S degrees was +4.44 and +2.91 kcal/mol, respectively) components. The implications of these finding are discussed.  相似文献   

11.
The relationship between the structure of the N-terminal sequence of transthyretin (TTR) and the binding of thyroid hormone was studied. A recombinant human TTR and two derivatives of Crocodylus porosus TTRs, one with the N-terminal sequence replaced by that of human TTR (human/crocTTR), the other with the N-terminal segment removed (truncated crocTTR), were synthesized in Pichia pastoris. Subunit mass, native molecular weight, tetramer formation, cross-reactivity to TTR antibodies and binding to retinol-binding protein of these recombinant TTRs were similar to TTRs found in nature. Analysis of the binding affinity to thyroid hormones of recombinant human TTR showed a dissociation constant (Kd) for triiodothyronine (T3) of 53.26+/-3.97 nM and for thyroxine (T4) of 19.73+/-0.13 nM. These values are similar to those found for TTR purified from human serum, and gave a Kd T3/T4 ratio of 2.70. The affinity for T4 of human/crocTTR (Kd=22.75+/-1.89 nM) was higher than those of both human TTR and C. porosus TTR, but the affinity for T3 (Kd=5.40+/-0.25 nM) was similar to C. porosus TTR, giving a Kd T3/T4 ratio of 0.24. A similar affinity for both T3 (Kd=57.78+/-5.65 nM) and T4 (Kd=59.72+/-3.38 nM), with a Kd T3/T4 ratio of 0.97, was observed for truncated crocTTR. The obtained results strongly confirm the hypothesis that the unstructured N-terminal region of TTR critically influences the specificity and affinity of thyroid hormone binding to TTR.  相似文献   

12.
The thermodynamics governing the denaturation of RNA duplexes containing 8 bp and a central tandem mismatch or 10 bp were evaluated using UV absorbance melting curves. Each of the eight tandem mismatches that were examined had one U-U pair adjacent to another noncanonical base pair. They were examined in two different RNA duplex environments, one with the tandem mismatch closed by G.C base pairs and the other with G.C and A.U closing base pairs. The free energy increments (Delta Gdegrees(loop)) of the 2 x 2 loops were positive, and showed relatively small differences between the two closing base pair environments. Assuming temperature-independent enthalpy changes for the transitions, (Delta Gdegrees(loop)) for the 2 x 2 loops varied from 0.9 to 1.9 kcal/mol in 1 M Na(+) at 37 degrees C. Most values were within 0.8 kcal/mol of previously estimated values; however, a few sequences differed by 1.2-2.0 kcal/mol. Single strands employed to form the RNA duplexes exhibited small noncooperative absorbance increases with temperature or transitions indicative of partial self-complementary duplexes. One strand formed a partial self-complementary duplex that was more stable than the tandem mismatch duplexes it formed. Transitions of the RNA duplexes were analyzed using equations that included the coupled equilibrium of self-complementary duplex and non-self-complementary duplex denaturation. The average heat capacity change (DeltaC(p)) associated with the transitions of two RNA duplexes was estimated by plotting DeltaH degrees and DeltaS degrees evaluated at different strand concentrations as a function of T(m) and ln T(m), respectively. The average DeltaC(p) was 70 +/- 5 cal K(-)(1) (mol of base pairs)(-)(1). Consideration of this heat capacity change reduced the free energy of formation at 37 degrees C of the 10 bp control RNA duplexes by 0.3-0.6 kcal/mol, which may increase Delta Gdegrees(loop) values by similar amounts.  相似文献   

13.
14.
Rabbit anti-dinitrophenyl antibody from a serum pool was obtained as five fractions of purified specific antibody by a limiting antigen precipitation method. Each fraction had a different binding affinity for epsilon-N-2,4-dinitrophenyl-L-lysine. The free energy changes for hapten binding to the five antibody fractions varied from -8.35 to -10.0 kcal/mol. An average deltaH of -13.9 kcal/mol was measured for the fractions with a batch calorimeter. The results indicate no significant correlation between enthalpy changes and free energy changes. However, a statistically significant correlation between the free energy changes and the entropy changes was found. The enthalpy of the anti-dinitrophenyl antibody interaction with multivalent dinitrophenyl human serum albumin was determined. These are the first enthalpy measurements of an antibody antigen reaction in which the intrinsic binding enthalpy between the antibody and the determinant group is known. The deltaH for the antigen binding reaction was -10.1 kcal/mol which is 3.8 kcal/mol less exothermic than the deltaH for the hapten binding reaction. The interactions that could lead to such a difference in enthalpy are discussed.  相似文献   

15.
Conformational stability and mechanism of folding of ribonuclease T1   总被引:5,自引:0,他引:5  
Urea and thermal unfolding curves for ribonuclease T1 (RNase T1) were determined by measuring several different physical properties. In all cases, steep, single-step unfolding curves were observed. When these results were analyzed by assuming a two-state folding mechanism, the plots of fraction unfolded protein versus denaturant were coincident. The dependence of the free energy of unfolding, delta G (in kcal/mol), on urea concentration is given by delta G = 5.6 - 1.21 (urea). The parameters characterizing the thermodynamics of unfolding are: midpoint of the thermal unfolding curve, Tm = 48.1 degrees C, enthalpy change at Tm, delta Hm = 97 kcal/mol, and heat capacity change, delta Cp = 1650 cal/mol deg. A single kinetic phase was observed for both the folding and unfolding of RNase T1 in the transition and post-transition regions. However, two slow kinetic phases were observed during folding in the pre-transition region. These two slow phases account for about 90% of the observed amplitude, indicating that a faster kinetic phase is also present. The slow phases probably result from cis-trans isomerization at the 2 proline residues that have a cis configuration in folded RNase T1. These results suggest that RNase T1 folds by a highly cooperative mechanism with no structural intermediates once the proline residues have assumed their correct isomeric configuration. At 25 degrees C, the folded conformation is more stable than the unfolded conformations by 5.6 kcal/mol at pH 7 and by 8.9 kcal/mol at pH 5, which is the pH of maximum stability. At pH 7, the thermodynamic data indicate that the maximum conformational stability of 8.3 kcal/mol will occur at -6 degrees C.  相似文献   

16.
Jung HI  Cooper A  Perham RN 《Biochemistry》2002,41(33):10446-10453
Structural studies have shown that electrostatic interactions play a major part in the binding of dihydrolipoyl dehydrogenase (E3) to the peripheral subunit-binding domain (PSBD) of the dihydrolipoyl acyltransferase (E2) in the assembly of the pyruvate dehydrogenase multienzyme complex of Bacillus stearothermophilus. The binding is characterized by a small, unfavorable enthalpy change (deltaH degrees = +2.2 kcal/mol) and a large, positive entropy change (TdeltaS degrees = +14.8 kcal/mol). The contributions of individual surface residues of the PSBD of E2 to its interaction with E3 have been assessed by alanine-scanning mutagenesis, surface plasmon resonance detection, and isothermal titration calorimetry. The mutation R135A in the PSBD gave rise to a significant decrease (120-fold) in the binding affinity; two other mutations (R139A and R156A) were associated with smaller effects. The binding of the R135A mutant to E3 was accompanied by a favorable enthalpy (deltaH degrees = -2.6 kcal/mol) and a less positive entropy change (TdeltaS degrees = +7.2 kcal/mol). The midpoint melting temperature (T(m)) of E3-PSBD complexes was determined by differential scanning calorimetry. The R135A mutation caused a significant decrease (5 degrees C) in the T(m), compared with the wild-type complex. The results reveal the importance of Arg135 of the PSBD as a key residue in the molecular recognition of E3 by E2, and as a major participant in the overall entropy-driven binding process. Further, the effects of mutagenesis on the deltaCp of subunit association illustrate the difficulties in attributing changes in heat capacity to specific classes of interactions.  相似文献   

17.
L Richard  L Genberg  J Deak  H L Chiu  R J Miller 《Biochemistry》1992,31(44):10703-10715
Phase grating spectroscopy has been used to follow the optically triggered tertiary structural changes of carboxymyoglobin (MbCO) and carboxyhemoglobin (HbCO). Probe wavelength and temperature dependencies have shown that the grating signal arises from nonthermal density changes induced by the protein structural changes. The material displaced through the protein structural changes leads to the excitation of coherent acoustic modes of the surrounding water. The coupling of the structural changes to the fluid hydrodynamics demonstrates that a global change in the protein structure is occurring in less than 30 ps. The global relaxation is on the same time scale as the local changes in structure in the vicinity of the heme pocket. The observed dynamics for global relaxation and correspondence between the local and global structural changes provides evidence for the involvement of collective modes in the propagation of the initial tertiary conformational changes. The energetics can also be derived from the acoustic signal. For MbCO, the photodissociation process is endothermic by 21 +/- 2 kcal/mol, which corresponds closely to the expected Fe-CO bond enthalpy. In contrast, HbCO dissipates approximately 10 kcal/mol more energy relative to myoglobin during its initial tertiary structural relaxation. The difference in energetics indicates that significantly more energy is stored in the hemoglobin structure and is believed to be related to the quaternary structure of hemoglobin not present in the monomeric form of myoglobin. These findings provide new insight into the biomechanics of conformational changes in proteins and lend support to theoretical models invoking stored strain energy as the driving force for large amplitude correlated motions.  相似文献   

18.
Biophysical evidences suggest that transthyretin (TTR) tetramer dissociation to the monomeric intermediate and subsequent polymerization leads to amyloid fibril formation, which is implicated in the pathogenesis of familial amyloid polyneuropathy (FAP) and senile systemic amyloidosis (SSA). Hence, inhibition of fibril formation is considered a potential therapeutic strategy. Here in we demonstrate that curcumin, a phenolic constituent of curry spice turmeric, binds to the active site of TTR through fluorescence quenching and ANS displacement studies. Binding of curcumin appears to inhibit the denaturant induced tertiary and quaternary structural changes in TTR as monitored by intrinsic emission fluorescence and glutaraldehyde cross-linking studies. However, curcumin did not bind to TTR at acidic pH. Protonation/ isomerization of the side chain oxygen atoms of curcumin at low pH might hamper the binding. These results suggest that curcumin binds to and stabilizes TTR thereby highlight the importance of the side chain conformations of the ligand in binding to TTR.  相似文献   

19.
20.
Previous results (Brouwer and van den Berg, Toxicol. Appl. Pharmacol., 85 (1986) 301) indicated preferential binding of a hydroxylated metabolite of tetrachlorobiphenyl to transthyretin (TTR) a carrier of thyroxine (T4). In the present study it was investigated whether the T4 binding site of TTR could be occupied specifically by hydroxylated chlorinated aromatic compounds using chlorinated phenol congeners as model compounds in a competition assay with [125I]T4. Chlorinated aromatics such as 2,3-dichlorobenzene and 3,4,3',4'-tetrachlorobiphenyl, and phenols such as 4-hydroxybiphenyl and phenol were inefficient competitors. All chlorinated phenols tested were competitors for the T4 binding site of TTR. The ranking in competition was pentachlorophenol (PCP) greater than trichlorophenols greater than dichlorophenols greater than monochlorophenols. Structures with chlorine in both ortho positions to the hydroxyl group were more efficient competitors. The relative affinity of binding of pentachlorophenol (PCP) to TTR was about twice that of T4. Scatchard analysis showed that PCP mainly decreased the affinity constant K11 while the binding capacity R1 was not altered, indicating a competitive type of inhibition. PCP was also able to compete with T4 sites on albumin with a relative affinity of 0.25. T4 binding to thyroid binding globulin (TBG) was much less affected by interference of PCP (relative affinity 0.001). The results indicate a specific interaction of chlorophenols with the T4 binding site of TTR.  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