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 46 毫秒
1.
Three new Hb S variants containing beta87 Leu, Trp, or Asp instead of Thr were expressed in yeast in order to further define the role of the beta87 position in stability and polymerization of deoxy Hb S. Previous studies showed that hydrophobicity at beta85 Phe and beta88 Leu is critical for stabilization of hemoglobin. Results with the three Hb S beta87 variants, however, showed minimal differences in stability, suggesting that beta87 amino acid hydrophobicity is not critical for stabilization of hemoglobin. Polymerization properties of the variants in the deoxy form, however, were affected by the beta87 amino acid. Polymerization of Hb S beta87 Thr --> Leu and Hb S beta87 Thr --> Trp was preceded by a delay time like Hb S, while Hb S beta87 Thr --> Asp did not show a delay time. In addition, changes in time required for half polymer formation (T1/2) as a function of hemoglobin concentration for Hb S beta87 Thr --> Asp were similar to that for beta87 Thr --> Gln. Hb S beta87 Thr --> Leu polymerized at a lower hemoglobin concentration than Hb S while beta87 Thr --> Trp and Hb S beta87 Thr --> Asp required much higher hemoglobin concentrations for polymer formation. Critical concentration required for deoxy Hb S beta87 Thr --> Asp polymerization was 6- and 2.3-fold greater than that for Hb S beta85 Phe --> Glu and Hb S beta88 Leu --> Glu, respectively. These results suggest that even though beta87 Thr is not a direct interaction site for beta6 Val in deoxy Hb S polymers, it does play a critical role in formation of the hydrophobic acceptor pocket which then promotes protein-protein interactions facilitating formation of stable nuclei and polymers of deoxy Hb S.  相似文献   

2.
The Ser-beta 174 residue of the Escherichia coli H(+)-ATPase beta subunit has been shown to be near the catalytic site together with Gly-beta 149, Gly-beta 172, Glu-beta 192, and Val-beta 198 (Iwamoto, A., Park, M.-Y., Maeda, M., and Futai, M. (1993) J. Biol. Chem. 268, 3156-3160). In this study, we introduced various residues at position 174 and found that the larger the side chain volume of the residue introduced, the lower the enzyme activity became. The Phe-beta 174 mutant was defective in energy coupling between catalysis and transport, whereas the Leu-beta 174 mutant could couple efficiently, although both mutants had essentially the same ATPase activities (approximately 10% of the wild type). The defective energy coupling of the Phe-beta 174 mutant was suppressed by the second mutation (Arg-alpha 296-->Cys) in the alpha subunit. The Cys-alpha 296/Phe-beta 174 mutant had essentially the same membrane ATPase activity as the Phe-beta 174 single mutant when assayed under the conditions that stabilize the double mutant enzyme. These results indicate the importance of the alpha/beta interaction, especially that between the regions near Arg-alpha 296 and Ser-beta 174, for energy coupling in the H(+)-ATPase. The 2 residues (Ser-beta 174 and Arg-alpha 296) may be located nearby at the interface of the two subunits. About 1 mol of N-[14C]ethylmaleimide could bind to 1 mol of the alpha subunit of Cys-alpha 296/Phe-beta 174 or Cys-alpha 296 mutant ATPase, but could not inhibit the enzyme activity. This is the first intersubunit mutation/suppression approach to ATPase catalysis and its energy coupling.  相似文献   

3.
To provide quantitative information on the sites that promote polymerization of sickle hemoglobin (HbS) after formation of the initial hydrophobic bond involving Val-6(beta) [E6V(beta)] and also to provide hemoglobins with an enhanced polymerization that could be used in a mouse model for sickle cell anemia, we have expressed recombinant double, triple, and quadruple HbS mutants with substitutions on both the alpha- and beta-chains, E6V(beta)/E121R(beta), D75Y(alpha)/E6V(beta)/E121R(beta) and D6A(alpha)/D75Y(alpha)/E6V(beta)/E121R(beta). These recombinant hemoglobins were extensively characterized by high-performance liquid chromatography analysis, sodium dodecyl sulfate-polyacrylamide gel electrophoresis, isoelectric focusing, amino acid analysis, and mass spectroscopy. They retained the functional properties of the Hb tetramer and polymerized in a linear manner at progressively lower Hb concentration as a function of the degree of substitution, suggesting that these remote sites (alphaD6A, alphaD75Y, and betaE121R) on the alpha- and beta-chains exhibit additive, enhanced polymerization properties. The quadruple mutant has a polymerization concentration close to that of the purified SAD hemoglobin from transgenic mouse red blood cells consisting of HbS, Hb Antilles, and Hb D-Punjab. Normal mouse Hb increases the polymerization concentration of each mutant. Thus, the general approach of using recombinant Hbs as described here should prove useful in elucidating the quantitative aspects of the mechanism of HbS polymerization and in identifying the contribution of individual sites to the overall process. The strategy described here demonstrates the feasibility of a systematic approach to achieve future recombinant HbS mutants that could provide a new generation of the transgenic mouse model for sickle cell anemia.  相似文献   

4.
We have examined the effect of 2, 3-diphosphoglycerate (DPG) on the solubility of deoxy-sickle hemoglobin (deoxy-Hb S) under conditions such that concentration, pH, and osmolarity of deoxy-Hb S solutions approached physiological. The range of DPG/Hb molar ratios encompassed the extremes found for this ratio in erythrocytes from individuals with sickle cell anemia. After monomer-polyer equilibrium had been established, the phases were separated by centrifugation and assayed for concentrations of Hb and DPG. DPG had no effect on the solubility of deoxy-Hb S. Furthermore, at DPG/Hb molar ratios less than one, there was no preferential incorporation of deoxy-Hb S containing bound DPG into polymers. At DPG/Hb molar ratios greater than one, concentrations of free DPG in monomer and polymer phases were virtually identical. Thus, under the specified equilibrium conditions, DPG is not a determining factor in the polymerization of deoxy-Hb S.  相似文献   

5.
The in-vitro synthesis of hemoglobin (Hb) chains was studied among 60 Hb S heterozygotes (AS) having different quantities of Hb S, including five with an associated alpha-chain heterozygosity (ASAG). Hematologic values and hemoglobin composition were studied in these cases and in 15 other ASAG heterozygotes. The percentages of Hb S (which fell between 27% and 42%) and the mean corpuscular volume values correlated directly with the alpha/non-alpha values, confirming previous suggestions (Huisman, Hemoglobin 1:349, 1977) that the concomitant occurrence of an alpha-thalassemia-2 heterozygosity (alpha alpha(0)/alpha alpha; beta/beta(S)) or homozygosity (alpha(0) alpha/alpha(0) alpha; beta/beta(S)) resulted in intermediate or lower levels of Hb S compared with Hb S heterozygotes having four active alpha-chain genes (alpha alpha/alpha alpha; beta/beta(S)). Among ASAG heterozygotes, the occurrence of low (about 25%), intermediate (about 33%), or high (about 45%) proportions of an alpha-chain variant resulting from a variability in the number of active alpha-chain genes due to alpha-thal-2 coincided with high (39%), intermediate (34%), or low (28%) levels of Hb S, respectively. However, the overlap of biosynthetic data between Hb S heterozygotes with four, three, or two active alpha-chain genes prevents a reliable diagnosis in individual cases.  相似文献   

6.
To provide new insights into ligand-A1 adenosine receptor (A1AR) interactions, site-directed mutagenesis was used to test the role of several residues in the first four transmembrane domains of the human A1AR. First, we replaced eight unique A1AR residues with amino acids present at corresponding transmembrane (TM) positions of A2AARs. We also tested the role of carboxamide amino acids in TMs 1-4, and the roles of Val-87, Leu-88, and Thr-91 in TM3. Following conversion of Gly-14 in TM1 to Thr-14, the affinity for adenosine agonists increased 100-fold, and after Pro-25 in TM1 was converted to Leu-25, the affinity for agonists fell. After conversion of TM3 sites Thr-91 to Ala-91, and Gln-92 to Ala-92, the affinity for N6-substituted agonists was reduced, and binding of ligands without N6 substituents was eliminated. When Leu-88 was converted to Ala-88, the binding of ligands with N6 substituents was reduced to a greater extent than ligands without N6 substituents. Following conversion of Pro-86 to Phe-86, the affinity for N6-substituted agonists was lost, and the affinity for ligands without N6 substitution was reduced. These observations strongly suggest that Thr-91 and Gln-92 in TM3 interact with the adenosine adenine moiety, and Leu-88 and Pro-86 play roles in conferring specificity for A1AR selective compounds. Using computer modeling based on the structure of rhodopsin, a revised model of adenosine-A1AR interactions is proposed with the N6-adenine position oriented toward the top of TM3 and the ribose group interacting with the bottom half of TMs 3 and 7.  相似文献   

7.
The determinants of insulin-like growth factor (IGF) binding to its binding proteins (IGFBPs) are poorly characterized in terms of important residues in the IGFBP molecule. We have previously used tyrosine iodination to implicate Tyr-60 in the IGF-binding site of bovine IGFBP-2 (Hobba, G. D., Forbes, B. E., Parkinson, E. J., Francis, G. L., and Wallace, J. C. (1996) J. Biol. Chem. 271, 30529-30536). In this report, we show that the mutagenic replacement of Tyr-60 with either Ala or Phe reduced the affinity of bIGFBP-2 for IGF-I (4.0- and 8.4-fold, respectively) and for IGF-II (3.5- and 4.0-fold, respectively). Although adjacent residues Val-59, Thr-61, Pro-62, and Arg-63 are well conserved in IGFBP family members, Ala substitution for these residues did not reduce the IGF affinity of bIGFBP-2. Kinetic analysis of the bIGFBP-2 mutants on IGF biosensor chips in the BIAcore instrument revealed that Tyr-60 --> Phe bIGFBP-2 bound to the IGF-I surface 3.0-fold more slowly than bIGFBP-2 and was released 2.6-fold more rapidly than bIGFBP-2. We therefore propose that the hydroxyl group of Tyr-60 participates in a hydrogen bond that is important for the initial complex formation with IGF-I and the stabilization of this complex. In contrast, Tyr-60 --> Ala bIGFBP-2 associated with the IGF-I surface 5.0-fold more rapidly than bIGFBP-2 but exhibited an 18.4-fold more rapid release from this surface compared with bIGFBP-2. Thus both the aromatic nature and the hydrogen bonding potential of the tyrosyl side chain of Tyr-60 are important structural determinants of the IGF-binding site of bIGFBP-2.  相似文献   

8.
Recent studies noted the contribution of alpha42Tyr to the T-R-dependent UV resonance Raman (UVRR) spectral changes of HbA [Nagai, M., et al. (1996) J. Mol. Struct. 379, 65-75; Huang, S., et al. (1997) Biochemistry 36, 6197-6206], but the observed UVRR changes of the Tyr residue cannot be fully interpreted with alpha42Tyr alone. To identify the remaining contributions, the 235 nm-excited UVRR spectra of Tyr mutant Hbs at alpha140, beta35, and beta145 were investigated here. The Fe-His stretching mode demonstrated that all of these mutant Hbs take the T structure in the deoxy form under these experimental conditions. The UVRR change of the Trp residue of these mutants upon the T-R transition was the same as that in HbA, indicating that the T-R-dependent UVRR change of beta37Trp is not due to stacking with Tyr residues but is due to the formation or destruction of a hydrogen bond. The recombinant Hbs beta35Tyr --> Phe and beta35Tyr --> Thr both exhibited UVRR spectra identical with that of HbA, meaning that beta35Tyr is not responsible. In the spectra of des(beta146His,beta145Tyr)Hb with inositol hexaphosphate, the frequency shift of the Tyr RR bands was the same as that in HbA but the intensity enhancement in the CO form was small, suggesting that beta145Tyr contributes to a part of the intensity change, but scarcely relates to the frequency shift. In the spectra of Hb Rouen (alpha140Tyr --> His), the frequency shifts of bands at 1617 (Y8a) and 1177 (Y9a) cm-1 following ligation were half of those in HbA, while the intensity enhancement was not detected. This result means that alpha140Tyr is responsible for both the frequency shift and the intensity changes. It is suggested that the frequency shift of the Tyr RR bands upon the T --> R transition is due to changes in the hydrogen bonding state of alpha42- and alpha140Tyr and that the intensity enhancement is due to changes in the environment of the penultimate Tyr in both alpha and beta subunits (alpha140 and beta145). These alterations in the vibrational spectra clearly demonstrate which tyrosine residues are involved in the T-R transition as a result of modification of their local environments.  相似文献   

9.
The 20S proteasome from the methanoarchaeon Methanosarcina thermophila was produced in Escherichia coli and characterized. The biochemical properties revealed novel features of the archaeal 20S proteasome. A fully active 20S proteasome could be assembled in vitro with purified native alpha ring structures and beta prosubunits independently produced in Escherichia coli, which demonstrated that accessory proteins are not essential for processing of the beta prosubunits or assembly of the 20S proteasome. A protein complex with a molecular mass intermediate to those of the alpha7 ring and the 20S proteasome was detected, suggesting that the 20S proteasome is assembled from precursor complexes. The heterologously produced M. thermophila 20S proteasome predominately catalyzed cleavage of peptide bonds carboxyl to the acidic residue Glu (postglutamyl activity) and the hydrophobic residues Phe and Tyr (chymotrypsinlike activity) in short chromogenic and fluorogenic peptides. Low-level hydrolyzing activities were also detected carboxyl to the acidic residue Asp and the basic residue Arg (trypsinlike activity). Sodium dodecyl sulfate and divalent or monovalent ions stimulated chymotrypsinlike activity and inhibited postglutamyl activity, whereas ATP stimulated postglutamyl activity but had little effect on the chymotrypsinlike activity. The results suggest that the 20S proteasome is a flexible protein which adjusts to binding of substrates. The 20S proteasome also hydrolyzed large proteins. Replacement of the nucleophilic Thr1 residue with an Ala in the beta subunit abolished all activities, which suggests that only one active site is responsible for the multisubstrate activity. Replacement of beta subunit active-site Lys33 with Arg reduced all activities, which further supports the existence of one catalytic site; however, this result also suggests a role for Lys33 in polarization of the Thr1 N, which serves to strip a proton from the active-site Thr1 Ogamma nucleophile. Replacement of Asp51 with Asn had no significant effect on trypsinlike activity, enhanced postglutamyl and trypsinlike activities, and only partially reduced lysozyme-hydrolyzing activity, which suggested that this residue is not essential for multisubstrate activity.  相似文献   

10.
To develop an assay for hemin dissociation, His64(E7) was replaced by Tyr in sperm whale myoglobin producing a holoprotein with a distinct green color due to an intense absorption band at 600 nm. Val68(E11) was replaced by Phe in the same protein to increase its stability. When excess Tyr64-Val68 apoglobin is mixed with either metmyoglobin or methemoglobin, the solution turns from brown to green, and the absorbance changes can be used to measure complete time courses for hemin dissociation from either holoprotein. This assay has been used to measure rates of hemin dissociation from native metmyoglobin, four myoglobin mutants (Ala64(E7), Ala68(E11), Phe68(E11), and Glu45(CD3)), native methemoglobin, valence hybrid hemoglobins, and two mutant hemoglobins ((alpha(Gly-E7)beta(native))2, and (alpha(native)beta(Gly-E7))2). Two kinetic phases were observed for hemin dissociation from native human hemoglobin at pH 7.0 and 37 degrees C. Valence and mutant hybrid hemoglobins were used to assign the faster phase (k = 7.8 +/- 2.0 h-1) to hemin dissociation from ferric beta subunits and the slower (k = 0.6 +/- 0.15 h-1) to dissociation from alpha subunits. The corresponding rate for wild-type metmyoglobin is 0.007 +/- 0.004 h-1.  相似文献   

11.
During the past decade new procedures have been developed for the isolation of RNA from a few mL of freshly collected blood. This material is reverse transcribed and the resulting cDNA can be used for the determination of the ratios between different types of globin mRNA, namely alpha 2/alpha 1, alpha/zeta, alpha/beta, gamma/beta, beta A/beta X, delta beta Lep/beta, and G gamma/A gamma. Details about these polymerase chain reaction-based methods are reviewed, and information about their usefulness in studying alpha-thalassemia, beta-thalassemia, sickle cell anemia and other beta-globin gene abnormalities, Hb Lepore heterozygosity, and heterozygosity for alpha 2- or alpha 1-globin gene mutations will be provided. The methods are also most useful in characterizing the mRNA types in single, in vitro cultured, BFU-E colonies; in colonies derived from cells of a Hb S heterozygote; for instance, the beta A- and beta(S)-mRNAs were present in all colonies and in about equal quantities, while many of those cells from a subject with a somatic cell mutant (Hb Costa Rica) contained beta A-mRNA and no beta-Costa Rica mRNA, and only a few had both types. The techniques described have considerable diagnostic value and offer a rather simple approach to the study of some of the listed diseases.  相似文献   

12.
To study the interaction of T cell receptor with its ligand, a complex of a major histocompatibility complex molecule and a peptide, we derived H-2Kd-restricted cytolytic T lymphocyte clones from mice immunized with a Plasmodium berghei circumsporozoite peptide (PbCS) 252-260 (SYIPSAEKI) derivative containing photoreactive Nepsilon-[4-azidobenzoyl] lysine in place of Pro-255. This residue and Lys-259 were essential parts of the epitope recognized by these clones. Most of the clones expressed BV1S1A1 encoded beta chains along with specific complementary determining region (CDR) 3beta regions but diverse alpha chain sequences. Surprisingly, all T cell receptors were preferentially photoaffinity labeled on the alpha chain. For a representative T cell receptor, the photoaffinity labeled site was located in the Valpha C-strand. Computer modeling suggested the presence of a hydrophobic pocket, which is formed by parts of the Valpha/Jalpha C-, F-, and G-strands and adjacent CDR3alpha residues and structured to be able to avidly bind the photoreactive ligand side chain. We previously found that a T cell receptor specific for a PbCS peptide derivative containing this photoreactive side chain in position 259 similarly used a hydrophobic pocket located between the junctional CDR3 loops. We propose that this nonpolar domain in these locations allow T cell receptors to avidly and specifically bind epitopes containing non-peptidic side chains.  相似文献   

13.
14.
Hybridization experiments between normal Hb tetramers (Fe2+ Hb) and those with four metal-substituted hemes (i.e., replacement of Fe2+ by Co2+, Mg2+, Mn2+, Mn3+, Ni2+, or Zn2+) have revealed unexpected behavior. These homometallic Hbs have previously served as models that mimic the deoxy or oxy properties of normal Fe2+ Hb. In this study, hybrids were composed of one alpha 1 beta 1 dimer that is metal-substituted at both hemes, in association with a second dimer alpha 2 beta 2 that has normal Fe2+ hemes. Both metal-substituted subunits are unligated, whereas the two Fe2+ subunits either are both unligated or both ligated with O2, CO, or CN. It was found that four of the metal-substituted Hbs (Mg2+ Hb, Mn2+ Hb, Ni2+ Hb, and Zn2+ Hb) did not form detectable amounts of heterometallic hybrids with normal Fe2+ Hb even though (i) their homometallic parents formed tight tetrameric complexes with stabilities similar to that of Fe2+ Hb and (ii) hybrids with metal substitution at both alpha sites or both beta sites are known to form readily. This striking positional effect was independent of whether the normal Fe2+ hemes were ligated and of which ligand was used. These findings indicate that surprisingly large changes in tetramer behavior can arise from small and subtle perturbations at the heme sites. Possible origins of these effects are considered.  相似文献   

15.
To delineate the phospholipase C (PLC; EC 3.1.4.3) beta2 sequences involved in interactions with the beta-gamma subunits of G proteins, we prepared a number of mammalian expression plasmids encoding a series of PLC beta2 segments that span the region from the beginning of the X box to the end of the Y box. We found the sequence extending from residue Glu-435 to residue Val-641 inhibited Gbeta-gamma-mediated activation of PLC beta2 in transfected COS-7 cells. This PLC beta2 sequence also inhibited ligand-induced activation of PLC in COS-7 cells cotransfected with cDNAs encoding the complement component C5a receptor and PLC beta2 but not in cells transfected with the alpha1B-adrenergic receptor, suggesting that the PLC beta2 residues (Glu-435 to Val-641) inhibit the Gbeta-gamma-mediated but not the Galpha-mediated effect. The inhibitory effect on Gbeta-gamma-mediated activation of PLC beta2 may be the result of the interaction between Gbeta-gamma and the PLC beta2 fragment. This idea was confirmed by the observation that a fusion protein comprising these residues (Glu-435 to Val-641) of PLC beta2 and glutathione S-transferase (GST) bound to Gbeta-gamma in an in vitro binding assay. The Gbeta-gamma-binding region was further narrowed down to 62 amino acids (residues Leu-580 to Val-641) by testing fusion proteins comprising various PLC beta2 sequences and GST in the in vitro binding assay.  相似文献   

16.
The Tyr alpha 42-Asp beta 99 intersubunit H-bond stabilizes the T quaternary structure in hemoglobin (Hb) tetramers. We had proposed that Tyr alpha 42 acts as an acceptor in this H-bond, because the tyrosine Y8a/8b and Y7a' UVRR (ultraviolet resonance Raman) bands shift in directions opposite to those expected if tyrosine is an H-bond donor. If Asp beta 99 is the H-bond donor, then it must be protonated in the T state, and would be a previously unrecognized contributor to the Bohr effect. This implication was strengthened by the discovery that an R-minus-T difference FTIR (Fourier transform infrared) band at 1693 cm-1, which might be a signal from protonated carboxylate, is missing in Hb Kempsey, a mutant in which Asp beta 99 is replaced by Asn. However, we now find that this FTIR signal is insensitive to 13C-labeling of the aspartate residues in Hb, and cannot arise from protonated Asp beta 99. There are no other difference signals in the 1700 cm-1 region at a sensitivity of one COOH group. We conclude that Asp beta 99 is not protonated, and that the anomalous UVRR shifts must arise from compensating polarization of the Tyr alpha 42 OH. Candidates for this compensation are the H-bond donated by the Asp beta 94 backbone NH, and the nearby positive charge of Arg beta 40.  相似文献   

17.
Arg-Arg-Glu-Thr-Ala-Trp-Ala (RRETAWA) is a novel ligand peptide for integrin alpha5 beta1, which blocks alpha5 beta1-mediated cell adhesion to fibronectin (Koivunen, E., Wang, B., and Ruoslahti, E. (1994) J. Cell Biol. 124, 373-380). Here we have localized the binding site for RRETAWA on alpha5 beta1 using inhibitory monoclonal antibodies (mAbs) and site-directed mutagenesis. A cyclic peptide containing this sequence (*CRRETAWAC*) had little effect on the binding of most anti-alpha5 and anti-beta1 mAbs to alpha5 beta1 but completely blocked binding of the anti-alpha5 mAb 16 in a directly competitive manner. Hence, the binding site of RRETAWA appears to closely overlap with the epitope of mAb 16. *CRRETAWAC* also acted as a direct competitive inhibitor of the binding of Arg-Gly-Asp (RGD)-containing fibronectin fragments to alpha5 beta1, suggesting that the binding site for RRETAWA is also closely overlapping with that for RGD. However, differences between the binding sites of RRETAWA and RGD were apparent in that (i) RGD peptides allosterically inhibited the binding of mAb 16 to alpha5 beta1, and (ii) several mAbs that perturbed binding of alpha5 beta1 to RGD had little effect on binding of alpha5 beta1 to RRETAWA. A double mutation in alpha5 (S156G/W157S) blocked the interaction of both RRETAWA and mAb 16 with alpha5 beta1 but had no effect on fibronectin binding or on the binding of other anti-alpha5 mAbs. Ser156-Trp157 is located near the apex of a putative loop region on the upper surface of a predicted beta-propeller structure formed by the NH2-terminal repeats of alpha5. Our findings suggest that this sequence forms part of the ligand-binding pocket of alpha5 beta1. Furthermore, as Ser156-Trp157 is unique to the alpha5 subunit, it may be responsible for the specific recognition of RRETAWA by alpha5 beta1.  相似文献   

18.
The bivalve mollusc Lucina pectinata harbors sulfide-oxidizing chemoautotrophic bacteria and expresses a monomeric hemoglobin I, HbI, with normal O2, but extraordinarily high sulfide affinity. The crystal structure of aquomet Lucina HbI has revealed an active site with three residues not commonly found in vertebrate globins: Phe(B10), Gln(E7), and Phe(E11) (Rizzi, M., Wittenberg, J. B., Coda, A., Fasano, M., Ascenzi, P., and Bolognesi, M. (1994) J. Mol. Biol. 244, 86-89). Engineering these three residues into sperm whale myoglobin results in a triple mutant with approximately 700-fold higher sulfide affinity than for wild-type. The single crystal x-ray structure of the aquomet derivative of the myoglobin triple mutant and the solution 1H NMR active site structures of the cyanomet derivatives of both the myoglobin mutant and Lucina HbI have been determined to examine further the structural origin of their unusually high sulfide affinities. The major differences in the distal pocket is that in the aquomet form the carbonyl of Gln64(E7) serves as a H-bond acceptor, whereas in the cyanomet form the amido group acts as H-bond donor to the bound ligand. Phe68(E11) is rotated approximately 90 degrees about chi2 and located approximately 1-2 A closer to the iron atom in the myoglobin triple mutant relative to its conformation in Lucina HbI. The change in orientation potentially eliminates the stabilizing interaction with sulfide and, together with the decrease in size of the distal pocket, accounts for the 7-fold lower sulfide affinity of the myoglobin mutant compared with that of Lucina HbI.  相似文献   

19.
BACKGROUND: 3-Isopropylmalate dehydrogenase (IPMDH) and isocitrate dehydrogenase (ICDH) belong to a unique family of bifunctional decarboxylating dehydrogenases. Although the ICDH dimer catalyzes its reaction under a closed conformation, known structures of the IPMDH dimer (without substrate) adopt a fully open or a partially closed form. Considering the similarity in the catalytic mechanism, the IPMDH dimer must be in a fully closed conformation during the reaction. A large conformational change should therefore occur upon substrate binding. RESULTS: We have determined the crystal structure of IPMDH from Thiobacillus ferrooxidans (Tf) complexed with 3-isopropylmalate (IPM) at 2.0 A resolution by the molecular replacement method. The structure shows a fully closed conformation and the substrate-binding site is quite similar to that of ICDH except for a region around the gamma-isopropyl group. The gamma group is recognized by a unique hydrophobic pocket, which includes Glu88, Leu91 and Leu92 from subunit 1 and Val193' from subunit 2. CONCLUSIONS: A large movement of domain 1 is induced by substrate binding, which results in the formation of the hydrophobic pocket for the gamma-isopropyl moiety of IPM. A glutamic acid in domain 1, Glu88, participates in the formation of the hydrophobic pocket. The C beta and C gamma atoms of Glu88 interact with the gamma-isopropyl moiety of IPM and are central to the recognition of substrate. The acidic tip of Glu88 is likely to interact with the nicotinamide mononucleotide (NMN) ribose of NAD+ in the ternary complex. This structure clearly explains the substrate specificity of IPMDH.  相似文献   

20.
Aromatic amino acid aminotransferase (AroAT) and aspartate aminotransferase (AspAT) are known as dual-substrate enzymes, which can bind acidic and hydrophobic substrates in the same pocket (Kawaguchi, S., Nobe, Y., Yasuoka, J., Wakamiya, T., Kusumoto, S., and Kuramitsu, S. (1997) J. Biochem. (Tokyo) 122, 55-63). In order to elucidate the mechanism of hydrophobic substrate recognition, kinetic and thermodynamic analyses using substrates with different hydrophobicities were performed. They revealed that 1) amino acid substrate specificity (kmax/Kd) depended on the affinity for the substrate (1/Kd) and 2) binding of the hydrophobic side chain was enthalpy-driven, suggesting that van der Waals interactions between the substrate-binding pocket and hydrophobic substrate predominated. Three-dimensional structures of AspAT and AroAT bound to alpha-aminoheptanoic acid were built using the homology modeling method. A molecular dynamic simulation study suggested that the outward-facing position of the Arg292 side chain was the preferred state to a greater extent in AroAT than AspAT, which would make the hydrophobic substrate bound state of the former more stable. Furthermore, AroAT appeared to have a more flexible conformation than AspAT. Such flexibility would be expected to reduce the energetic cost of conformational rearrangement induced by substrate binding. These two mechanisms (positional preference of Arg and flexible conformation) may account for the high activity of AroAT toward hydrophobic substrates.  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