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.
The conformation of human placental alkaline phosphatase (EC 3.1.3.1) has been studied using the spectroscopic structural probes of pH difference spectroscopy, solvent perturbation difference spectroscopy, and circular dichroism. Of the 37 ± 1 tyrosine residues in placental alkaline phosphatase (PAP), 5 ± 1 residues are observed by pH difference spectroscopy to be “free” and presumed to be located on the surface of the enzyme molecule. The ionization of these 5 “free” tyrosyl groups is not time dependent and is reversible with a pKapp of 10.29. The remaining 32 ± 1 tyrosines are considered “buried” and ionization is observed to be both time dependent and irreversible. Treatment of the enzyme with 4 m guanidine-hydrochloride normalizes all 37 ± 1 tyrosine residues (pKapp = 10.08). The difference pH titration studies thus provide spectrophotometric evidence for a change in molecular conformation of PAP in the pH region of 10.5. Using solvent perturbation difference spectroscopy and circular dichroism, the local environments of tyrosine and tryptophan residues were elucidated for the native enzyme and the enzyme in the presence of ligands that influence catalytic function: inorganic phosphate (competitive inhibitor), l-phenylalanine (uncompetitive inhibitor), d-phenylalanine (noninhibitor). and Mg2+ ion (activator). The spectral observations from these studies led to the following interpretations: (i) the binding of inorganic phosphate, a competitive inhibitor, induces a conformational change in the enzyme that may alter the active site and thereby decrease enzyme catalytic function; (ii) perturbation with l-phenylalanine gives spectral results indicating a conformational change consistent with the postulate that this uncompetitive inhibitor prevents the dissociation of the phosphoryl enzyme intermediate; and (iii) Mg2+ ion causes a slight separation of the enzyme subunits, which could increase accessibility to the active site and, thus, enzyme activity.  相似文献   

2.
K Kuwajima  Y Ogawa  S Sugai 《Biochemistry》1979,18(5):878-882
A stopped-flow technique has been developed for the zero-time spectrophotometric titration of tyrosine residues in the purely native or in the purely alkaline denatured state of alpha-lactalbumin that undergoes an alkaline conformational transition in the pH region of tyrosine ionization. The progressive absorption change at 298 nm caused by a pH jump from neutral pH is shown to result from the change in ionization of the tyrosine residues brought about by a first-order process of the conformational transition. Extrapolation to zero time gives the titration curve for purely native alpha-lactalbumin. Similarly, the pH jump from highly alkaline pH gives the titration curve for the purely alkaline denatured protein. The method should be generally applicable to other proteins that contain tyrosines. Analysis of the titration curves suggests that the four tyrosines in native alpha-lactalbumin have pK values of 10.5, 11.8, 11.8, and 12.7, respectively. After the alkaline transconformation, all of them become titrated normally with a pK value of 10.3. A comparison of these results with the ionization behavior of tyrosines in hen egg white and human lysozymes is presented and discussed in terms of differences in the sequences of the proteins.  相似文献   

3.
Of the three tyrosine residues available for nitration by tetranitromethane in hemerythrin, nitration of tyrosine residue 70 has no effect on dissociation of octomers to monomers, but nitration of tyrosines 18 and/or 67 results in dissociation to monomers. The latter data suggests these residues are important for subunit association. The reactive sulfhydryl, the modification of which produces dissociation, was protected as a mixed disulfide during the nitration but was regenerated for analysis of the state of association. Residue 70 can be selectively modified because of its exposed position and perhaps because of its slightly lower pk of 6.9, compared to 7.3 as an average of all nitrotyrosines in a completely nitrated hemerythrin. Solvent perturbation studies in 20% Me2SO indicate that 3 tyrosines, in agreement with the nitration results, and 2 tryptophan residues are exposed; however, oxidation at a 2-fold molar excess of N-bromosuccinimide oxidizes three tryptophan whereas a 3.5-fold excess oxidizes all four, but results in a rapid active site destruction. Photo-oxidation with methylene blue results in oxidation of only two tryptophan residues. These data have been interpreted to indicate that two tryptophans are free and two are involved in subunit association. Photo-oxidation with methylene blue results in the destruction of three histidines but no decrease in active site absorption. Histidine modification with diethyloxydiformate shows that three histidines react with no change in active site absorption. These results indicate that four histidines are unreactive toward these modifying agents and are therefore either buried or are ligands to the iron.  相似文献   

4.
The ionization of the phenolic hydroxyl groups in calotropins DI and DII isolated from the latex of Calotropis gigantea has been studied by spectrophotometric titration at 295 nm in the pH range 6–13.2. Of the 12 tyrosine residues of calotropin DI and 13 tyrosine residues of calotropin DII, only four residues were ionized reversibly in the pH range 8.9–10.7 with the apparent pK of 9.7. The remaining tyrosine residues were ionized irreversibly in the pH range 11.2–13.2 with the apparent pK of 11.5. Both calotropins showed time-dependent ionization of phenolic groups at 295 nm in the pH range 11.5–12.0. Chemical modification with tetranitromethane suggested the presence of three tyrosine residues on the surface of each calotropin molecule.  相似文献   

5.
The reaction of hemoglobin with N-acetyl imidazole at neutral pH indicated that in carboxyhemoglobin 1.80 residues per heme were acetylated while in deoxyhemoglobin only 1.15 residues were available to the reagent. The reversible titration of these residues in alkali was followed by difference spectrophotometry at 245 nm. Hill plots of the titration data, assuming 2 residues titrable per heme an3 Δε = 10500 per tyrosyi residue upon ionization, showed a slope of 1.5 and a pH near 11. The average pK of these groups in carboxyhemoglobin was previously found to be near 10.5. Also. by difference spectrophotometry it was shown that exposure of deoxyhemoglobin to alkaline pH was accompanied by a modification of the Soret region of the absorption spectrum, which might indicate the appearance of liganded conformation in the deoxyhemoglobin system. The sedimentation velocity of deoxyhemoglobin demonstrated that at alkaline pH dissociation into duners occurred at pH's lower than 10, where no ionization of tyrosines was detectable. The titration of tyrosines was independent from protein concentration.The low availability of tyrosyl residues to acetylation in deoxyhemoglobin, the cooperativity of proton binling of these residues and the change in conformation of hemoglobin concomitant with their titration are all consistent with results of Simon et al., Moffat, and Moffat et al., and with the model proposed by Perutz for explaining the heme-heme interaction. The free energy of the pK shift of the tyrosyl residues in carboxy and deoxyhemoglobin can be included in the free energy of the heme-heme interaction.  相似文献   

6.
1. The dissociation of horse spleen apoferritin as a function of pH was analysed by sedimentation-velocity techniques. The oligomer is stable in the range pH2.8-10.6. Between pH2.8 and 1.6 and 10.6 and 13.0 both oligomer and subunits can be detected. At pH values between 1.6 and 1.0 the subunit is the only species observed, although below pH1.0 aggregation of the subunits to a particle sedimenting much faster than the oligomer occurs. 2. When apoferritin is first dissociated into subunits at low pH values and then dialysed into buffers of pH1.5-5.0, the subunit reassociates to oligomer in the pH range 3.1-4.3. 3. U.v.-difference spectroscopy was used to study conformational changes occurring during the dissociation process. The difference spectrum in acid can be accounted for by the transfer of four to five tyrosine residues/subunit from the interior of the protein into the solvent. This process is reversed on reassociation, but shows the same hysteresis as found by sedimentation techniques. The difference spectrum in alkali is more complex, but is consistent with the deprotonation of tyrosine residues, which appear to have rather high pK values. 4. In addition to the involvement of tyrosine residues in the conformational change at low pH values, spectral evidence is presented that one tryptophan residue/subunit also changes its environment before dissociation and subsequent to reassociation. 5. Analysis of the dissociation and reassociation of apoferritin at low pH values suggests that this is a co-operative process involving protonation and deprotonation of at least two carboxyl functions of rather low intrinsic pK. The dissociation at alkaline pH values does not appear to be co-operative. 6. Of the five tyrosine residues/subunit only one can be nitrated with tetranitromethane. Guanidination of lysine residues results in the modification of seven out of a total of nine residues/subunit. Nine out of the ten arginine residues/subunit react with cyclohexanedione.  相似文献   

7.
The chemical modification of cysteine and tyrosine residues in formyltetrahydrofolate synthetase from Clostridium thermoaceticum has been examined relative to enzymatic activity and reactivity of these groups in the native protein. 4,4′-Dipyridyl disulfide, dansylaziridine, and fluorescein mercuric acetate all reacted with just one of six sulfhydryls per enzyme subunit, resulting in activities of 100, 95 and 70%, respectively. The Km values for MgATP, formate, and tetrahydrofolate were unaltered in the modified enzymes. ATP did produce a 2.5-fold reduction in the rate of reaction between the enzyme and 4,4′-dipyridyl disulfide. Tetranitromethane reacted most rapidly with a single sulfhydryl group per subunit to produce a 20–30% loss in activity. Subsequent additions of tetranitromethane modified 2.2 tyrosines per subunit which was proportional to the loss of the remaining enzymatic activity. Folic acid, a competitive inhibitor, protected against modification of the tyrosines and the associated activity losses; however, the oxidation of the single sulfhydryl group and the initial 20–30% activity loss were unaffected. In the presence of folic acid, higher concentrations of tetranitromethane produced a loss of the remaining activity proportional to the modification of 1.2 tyrosines per subunit. It is proposed that at least 1 tyrosine critical for enzymatic activity is located at or near the folic acid/tetrahydrofolate binding site.  相似文献   

8.
J Wower  P Maly  M Zobawa  R Brimacombe 《Biochemistry》1983,22(10):2339-2346
The detailed surface topography of the Escherichia coli 30S ribosomal subunit has been investigated, with iodination catalyzed by immobilized lactoperoxidase as the surface probe. Under mild conditions, only proteins S3, S7, S9, S18, and S21 were iodinated to a significant and reproducible extent. These proteins were isolated from the iodinated subunits, and in each case, the individual tyrosine residues that had reacted were identified by standard protein sequencing techniques. The targets of iodination that could be positively established were as follows: in protein S3 (232 amino acids), the tyrosines at positions 167 and 192; in S7 (153 amino acids), tyrosines 84 and 152; in S9 (128 amino acids), tyrosine 89; in S18 (74 amino acids), tyrosine 3 (tentative); in S21 (70 amino acids), tyrosines 37 and 70. The results represent part of a broader program to investigate ribosomal topography at the amino acid-nucleotide level.  相似文献   

9.
Conformationally-linked dissociation equilibria of dimeric proteins have been examined to determine how experimentally obtainable parameters, such as the apparent dissociation constant, kD, and the apparent conformational transition constant, Kconf, are related to intrinsic subunit interaction constants, KA or KB, and intrinsic isomerization constants, K1 or k2. Limiting models are considered in which either the conformational change occurs before dissociation or in which the dissociation occurs before the conformational change, as well as a general model including both possibilities. Models are also considered in which three conformations are allowed or in which four subunits (tetramers) are involved. Simulated data for the dimer equilibria are presented to show how variation of protein concentration and variation of certain constants affect the proportion of various molecular species, the weight-average molecular weight, and the overall extent of conformational change. Methods are suggested to distinguish between the different limiting cases based upon the dependence of KD and/or Kconf on protein concentration, perturbant concentration, and temperature. It is concluded that methods used to calculate self-dissociation constants oligomeric proteins include linked isomerization reactions such that the equilibrium constant obtained should not be considered as a true subunit interaction term. Indeed, dissociation can occur under the influence of a perturbant with no change in the intrinsic affinity of the subunits but with the sole effect of the perturbant being on a linked conformational change. Additional experiments on the thermodynamics of the conformational change are required to determine the actual relationship. Depending on the complexity of the equilibria involved and the relative value of the equilibrium constants, the extent of the conformational change can vary directly with, vary inversely with, or he independent of the total protein concentration. Even when intrinsic subunit affinities are not affected by the perturbant, the extent of conformational change can vary with protein concentration. Interpretation of data from proteins which may be involved in conformationally-linked dissociation reactions, therefore, must be made with caution.  相似文献   

10.
The catalytic mechanism of 6-phosphogluconate dehydrogenase requires the inversion of a Lys/Glu couple from its natural ionization state. The pKa of these residues in free and substrate bound enzymes has been determined measuring by ITC the proton release/uptake induced by substrate binding at different pH values. Wt 6-phosphogluconate dehydrogenase from Trypanosoma brucei and two active site enzyme mutants, K185H and E192Q were investigated. Substrate binding was accompanied by proton release and was dependent on the ionization of a group with pKa 7.07 which was absent in the E192Q mutant. Kinetic data highlighted two pKa, 7.17 and 9.64, in the enzyme–substrate complex, the latter being absent in the E192Q mutant, suggesting that the substrate binding shifts Glu192 pKa from 7.07 to 9.64. A comparison of wt and E192Q mutant appears to show that the substrate binding shifts Lys185 pKa from 9.9 to 7.17. By comparing differences in proton release and the binding enthalpy of wt and mutant enzymes, the enthalpic cost of the change in the protonation state of Lys185 and Glu192 was estimated at ≈ 6.1 kcal/mol. The change in protonation state of Lys185 and Glu192 has little effect on Gibbs free energy, 240–325 cal/mol. However proton balance evidences the dissociation of other group(s) that can be collectively described by a single pKa shift from 9.1 to 7.54. This further change in ionization state of the enzyme causes an increase of free energy with a total cost of 1.2–2.3 kcal/mol to set the enzyme into a catalytically competent form.  相似文献   

11.
Abstract: Superoxide dismutase (SOD) catalyzes the nitration of specific tyrosine residues in proteins by peroxynitrite (ONOO), which may be the damaging gain-of-function resulting from mutations to SOD associated with familial amyotrophic lateral sclerosis (ALS). We found that disassembled neurofilament-L (light subunit) was more susceptible to tyrosine nitration catalyzed by SOD in vitro. Neurofilament-L was selectively nitrated compared with the majority of other proteins present in brain homogenates. Assembled neurofilament-L was more resistant to nitration, suggesting that the susceptible tyrosine residues were protected by intersubunit contacts in assembled neurofilaments. Electrospray mass spectrometry of trypsin-digested neurofilament-L showed that tyrosine 17 in the head region and tyrosines 138, 177, and 265 in α-helical coil regions of the rod domain of neurofilament-L were particularly susceptible to SOD-catalyzed nitration. Nitrated neurofilament-L inhibited the assembly of unmodified neurofilament subunits, suggesting that the affected tyrosines are located in regions important for intersubunit contacts. Neurofilaments are major structural proteins expressed in motor neurons and known to be important for their survival in vivo. We suggest that SOD-catalyzed nitration of neurofilament-L may have a significant role in the pathogenesis of ALS.  相似文献   

12.
The data on the pH dependence of the Km for Mg-ATP and the Vm of the ATPase of pig heart mitochondrial F1 indicate the presence of two groups of different pK's which modify the enzyme activity. The first pK at pH 9.6 ± 0.2 may be related to the possible presence of arginine and/or tyrosine residues in the ATPase site; the second pK at pH 7.2 ± 0.2 could be due to the presence of a histidine residue in the ATPase site or to the involvement of amino groups in the ATPase site. The inhibition induced by photooxidation in the presence of Rose Bengal is not pH dependent in the pH range corresponding to the pK of histidine. The inhibition induced by diethylpyrocarbonate cannot be reversed by hydroxylamine and the characteristics of this inhibition rather correspond to the reaction of the inhibitor with amino groups. Pyridoxal phosphate also inhibits the ATPase activity of F1 by reaction with amino groups. The presence of ATP or phosphate partially protects against the inhibition induced by diethylpyrocarbonate or pyridoxal phosphate, which indicates that amino groups may be directly or indirectly involved in the binding of nucleotide and phosphate to F1. Glutaraldehyde also inhibits the enzyme by reacting with amino groups and inducing a crosslinking of the subunits. The disappearance of subunit C is well correlated with the decrease of ATPase activity, indicating that subunit C is essential in the ATPase activity.  相似文献   

13.
The rapid reaction of diisopropylfluorophosphate with a tyrosine residue of human serum albumin at 0.02 m ionic strength involves prior rapid reversible binding characterized by a dissociation constant of 3.6 × 10?3m and an apparent pKa of 8.3. The rapid reaction of p-nitrophenyl acetate with human serum albumin (G. E. Means and M. L. Bender, 1975, Biochemistry14, 4989–4994) appears to involve the same tyrosine residue and is thus stoichiometrically inhibited by prior reaction with diisopropylfluorophosphate. Both reactions are strongly inhibited by decanoate anion, strongly retarded at higher ionic strength, and reflect strong rapidly reversible binding and abnormally low tyrosine pKa values. This reactive tyrosine residue thus appears to be located in a primary binding site for small apolar anions and to be closely associated with several cationic groups.  相似文献   

14.
The analysis of the effect of pH upon the rate of polymerization indicates that the activity of yeast RNA polymerase I is optimal between pH 7.5 and 9 and depends on the ionization state of two groups with apparent pKa values of 6.5 and 10. Yeast RNA polymerase I is extremely labile at acid pH. Below pH 5 the enzyme is irreversibly inactivated by [H+], with a second-order rate constant of 1.6 × 10?4m?1 min?1. Sucrose gradient sedimentation and gel electrophoresis analysis of the enzyme inactivated at acid pH indicates the sequential dissociation of several enzyme subunits. The polypeptides of 44,000 and 24,000 daltons dissociate first from the enzyme core followed by the dissociation of the polypeptides of 48,000 and 36,000 daltons.  相似文献   

15.
The reaction of tetranitromethane with human chorionic gonadotropin and its subunits has been investigated. The hormone consists of two subunits, α and β, containing four and three tyrosyl residues, respectively. Introduction of 1 nitrated tyrosine residue into the native hormone was accompanied by a 20% loss in immunological reactivity and a 50% loss in biological activity. This initial reaction occurred at α Tyr-88 and/or α Tyr-89. Exhaustive nitration of the hormone modified α tyrosines 65, 88, and 89 and resulted in 75% inactivation biologically and 50% immunologically. Either nitrated α subunit obtained by dissociation of the nitrated hormone recombined with the native β subunit to give a hormone whose activity was in reasonable agreement with that of the corresponding nitrated monomer. These results indicate involvement of α Tyr-88 and/or α Tyr 89 in binding of the hormone to its receptor. These residues are not required for binding to the β subunit, however. Tyr-65 of the α subunit is probably not involved in binding to either the β subunit or the hormone receptor. The β subunit obtained from the exhaustively nitrated hormone was unmodified and recombined with native α to give fully active hormone. About 25% of the protein was recovered as polymeric material following nitration; lesser amounts of crosslinked monomer were formed. Both were biologically inactive. The polymer products retained about 30% of the native immunological competence.Nitration of the isolated α subunit fully converted the remaining tyrosine (Tyr-37) to 3-nitrotyrosine in a two-step reaction. The fully nitrated α subunit did not recombine well with the native β subunit and the recombinant hormone has 10% or less of the native activity. Involvement of α Tyr-37 in binding to the β subunit is suggested by these data. However, exposure of this residue by a conformational change in the α subunit after dissociation of the native hormone, while it seems unlikely in view of the high disulfide content, is also consistent with the data. Reaction of the free β subunit with tetranitromethane resulted in complete nitration of Tyr-37, 85% nitration of Tyr-59, and 25% nitration of Tyr-82. The nitrated β subunit did not recombine well with native α but the isolated recombinant had two-thirds of the native activity. From these data we conclude that β Tyr-37 and/or β Tyr-59 are possibly involved in binding to the α subunit but do not have a role in the biological activity. Tyr-82 of β is apparently not involved in either subunit interactions or hormone-receptor binding.  相似文献   

16.
The fluorescence of 1,8-anilinonaphthalene sulfonate (ANS) was enhanced in the presence of ovine lutropin (oLH). Fluorescence titration curves were sigmoidal with 50% saturation between 200 and 500 μm. Exclusion chromatography experiments indicated that the hormone self-associates to form dimers in the presence of excess ANS. By contrast, the isolated a and β subunits of oLH caused a much smaller enhancement of the fluorescence of ANS and did not self-associate in its presence. Dissociation of the intact hormone into its subunits was accompanied by 1) a loss in the ability to enhance ANS fluorescence, 2) the appearance of a negative differential absorption spectrum whose magnitude indicated the increased solvent-exposure of at least two tyrosines, and 3) a loss in conformational rigidity as evidenced by a decrease in polarization (P) of tyrosyl fluorescence from ~0.17 to ~0.13. Similar rates of dissociation were obtained by all three measurements and the first order rate constant at pH 3.6 and 37 °C under conditions of low ionic strength was k = 0.18 min?1; at high ionic strength, e.g., 0.5 m KC1, dissociation was incomplete even after prolonged incubation. Acid-dissociated subunits recombined readily in 0.5 m acetate buffer, pH 5.3, and the recovery of the intrinsic absorption and fluorescence properties as well as the ability to enhance ANS fluorescence ranged between 70 and 90%. Titration of the isolated α and β subunits with acid or GdmCl had little or no effect on P, suggesting that residual secondary or tertiary structure is either absent, very stable, or its disruption does not alter the rigidity of the tyrosyl environment. The relatively high P for oLH-β (0.17) suggests a conformation which is rigid compared with oLH-α (0.13). P for both subunits decreased smoothly with increasing temperature between 20 and 70 °C. By contrast, oLH exhibited a thermal transition near 50 °C characterized by a drop in P from a value near that of β to a value near that of a as the subunits dissociated. Because α has more tyrosines with a higher average quantum yield, its fluorescence would be expected to dominate that of the hormone or of an equimolar mixture of subunits. Thus, most of the conformation changes which accompany dissociation and recombination appear to occur in the α subunit.  相似文献   

17.
The polymerization of aryl sulfatase A (aryl sulfate sulfohydrolase, EC 3.1.6.1) has been studied by frontal gel chromatography on Sephadex G-200 and Bio-Gel A-5m under various conditions of pH, ionic strength, and temperature. The aryl sulfatase A molecule exists as a monomer and as a dimer at pH 7.5 and pH 4.5, respectively. The extent of dissociation is markedly pH-, protein concentration-, and ionic strength-dependent. Only a small effect of temperature was observed. The enthalpy change (ΔHo) for the dissociation was ?2.5 ± 1 kcal/mol at pH 5.5–5.6, and the entropy change for dissociation of the enzyme dimer to two monomeric units was ?47 cal mol?1 deg?1. Sulfate ion has little effect on the extent of dissociation of the enzyme at pH 5.6. The present studies suggest that the dissociation of rabbit liver aryl sulfatase A is regulated by the ionization of amino acid residues whose apparent pK is between pH 5 and 6. The driving force for the association of the subunits of the enzyme is primarily ionic and/or ionic/hydrogen bond formation. The small enthalpy change and the fact that dissociation is strongly favored by an increase in the ionic strength suggest that hydrophobic interactions play only a minor role in stabilizing the dimeric quaternary structure relative to the monomeric state. The monomeric form of the enzyme exhibits the anomalous kinetics often observed with sulfatase A but the dimer does not show anomalous kinetics. Since aryl sulfatase A is probably in the dimeric form in the lysosome, the anomalous kinetics of the enzyme are unlikely to be of physiological importance in the intact lysosome.  相似文献   

18.
A preparation of ATPase from the membranes of Micrococcus lysodeikticus, solubilized and more than 95 %. pure, showed two main bands in analytical polyacrylamide gel electrophoresis. They did not correspond to isoenzymes because one band could be converted into the other by exposure to a mildly alkaline pH value. The conversion was paralleled by changes in molecular weight, circular dichroism and catalytic properties. Denaturation by pH at 25 °C was followed by means of circular dichroism, ultracentrifugation and polyacrylamide gel electrophoresis. A large conformational transition took place in the acid range with midpoints at about pH = 3.6 (I = 10?4 M), 4.3 (I = 0.03 M) and 5.3 (I = 0.1 M). The transition was irreversible. Strong aggregation of the protein occurred in this range of pH. The final product was largely random coil, but even at pH 1.5 dissociation into individual subunits was not complete. However, partial dissociation took place at pH 5 (I = 0.028 M). At this pH value the enzyme was inactive, but 20–30 % of the activity could be recovered when the pH was returned to 7.5.In the alkaline region the midpoint of the transition occurred near pH = 11 (I = 0.028 M). The pK of most of the tyrosine residues of the protein was about 10.9. The unfolding was irreversible and the protein was soon converted into peptide species with molecular weights lower than those determined for the subunits by gel clectrophoresis in the presence of sodium dodecyl sulphate. Conventional proteolysis did not account for the transformation.  相似文献   

19.
The fluorescence properties of the tyrosyl residues of human chorionic gonadotropin (hCG) and its α and β subunits have been examined. The effects of pH, guanidine, and disulfide cleavage on the intensity and polarization of the fluorescence suggest that the isolated subunits possess little, if any, tertiary structure beyond that which is stabilized by the disulfide bonds. Essentially all of the fluorescence of hCG and its subunits was accessible to quenching by iodide ions. Similar results were observed for several other proteins whose fluorescence originates from tyrosyl residues. Thus, we have confirmed and extended the conclusion of R. W. Cowgill ((1966) Biochim. Biophys. Acta120, 196) that the buried tyrosyl residues in ribonuclease fluoresce with a much lower quantum yield than those which are exposed. The dissociation of hCG into its subunits was accompanied by an increase in fluorescence, suggesting the exposure of tyrosyl residues. This was confirmed by difference absorption measurements which indicate a net exposure of two to three tyrosyl residues upon dissociation of the subunits. An additional 0.6 tyrosine was exposed when the disulfide bonds of the β-subunit were cleaved. The polarization of the fluorescence of hCG-β was high (P = 0.19) and, unlike several other proteins with high polarization, could not be lowered by denaturing conditions. Only by cleavage of the disulfide bonds could the fluorescence polarization of either subunit be lowered to a value (P = 0.08) characteristic of a random polypeptide. It appears that the disulfide bonds play an important role in maintaining the rigidity of the fluorescent tyrosyl residues, located at or near the surface of the protein.  相似文献   

20.
Porcine lutropin shares with ovine lutropin common structural features. They exhibit identical vacuum circular dichroism down to 170 nm with characteristic negative bands at 173, 194, and 210 nm. The band at 210 ± 1 nm is shifted to 201 nm upon dissociation with disappearance of the 194-nm band. For the two hormones the acid transition involves a significant loss of the three recognized periodic structures α helix, β sheet, and β turns of type II, unshields near the same number of tyrosyl residues (2.2 ± 0.7), and gives rise to an identical absorption difference at 287 nm in a two-step mechanism. However, pLH also exhibits noticeable differences: 5 to 10 times lower rates of acid transition with a lower pKa (3.7 ± 0.1) and different transition behavior of tyrosine residues compared to ovine lutropin.  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