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.
Hall J  Hall A  Pursifull N  Barbar E 《Biochemistry》2008,47(46):11940-11952
Dimerization of dynein light chain LC8 creates two symmetric grooves at the dimer interface with diverse binding capabilities. In addition to pH and protein concentration, dimerization is affected by phosphorylation, as illustrated by a phosphomimetic mutation that promotes dissociation of LC8 to a monomer and subsequent dissociation from the dynein complex in vitro. In this work we characterize the dynamic structure and unfolding profiles of an LC8 mutant, H55K, as a model for monomeric LC8 at neutral pH. Backbone (15)N relaxation experiments show that the monomer, while primarily ordered, has more heterogeneous dynamics relative to the LC8 dimer, predominantly in residues that ultimately form the binding groove, particularly those in beta 1 and beta 3 strands. This heterogeneity suggests that conformations that are primed for binding are sampled in the inactive monomer and favored in the active dimer. Further changes of LC8 backbone dynamics upon binding to short peptides from Swallow (Swa) and dynein intermediate chain (IC) were elucidated. The conformational heterogeneity apparent in the LC8 dimer is retained in LC8/IC but is lost in LC8/Swa, suggesting that the degree of ordering upon binding is ligand dependent. The reduced complexity of motion in LC8/Swa correlates with the less favorable entropy of binding of LC8 to Swa relative to IC. We propose that the conformational motility of beta 3 has functional significance in dimerization and in ligand binding. In the latter, beta 3 flexibility apparently accommodates different binding modes for different ligands resulting in ligand-specific conformational dynamics of the binding site that may impact other processes such as accessibility to phosphorylation.  相似文献   

2.
The 8-kDa light chain of dynein (DLC8) is ubiquitously expressed in various cell types. Other than serving as a light chain of the dynein complexes, this highly conserved protein has been shown to bind a larger number of proteins with diverse biological functions. DLC8 forms a homodimer via three-dimensional domain swapping of an internal beta-strand (the beta2-strand) at neutral pH. The protein undergoes non-reversible dimer-to-monomer dissociation when the pH value of the protein solution decreases. The three-dimensional structure of the DLC8 monomer determined by NMR spectroscopy at pH 3.0 showed that the protein is well folded. The major conformational change accompanied by dimer dissociation is in the beta2-strand of the protein, which undergoes transition from a beta-strand to a nascent alpha-helix. The monomer form of DLC8 is not capable of binding to target proteins. Insertion of two flexible amino acid residues in the tight beta1/beta2-loop dramatically stabilized the monomer conformation of the protein. NMR studies showed that the mutation altered the conformation as well as the three-dimensional domain swapping-mediated assembly of the DLC8 dimer. The mutant DLC8 was unable to bind to its targets even at physiological pH. The three-dimensional structure of the mutant protein in its monomeric form provides the structural basis of the mutation-induced stabilization of the monomer conformation. Based on the experimental data, we conclude that the formation of the beta2-strand swapping-mediated dimer is mandatory for the structure and function of DLC8. We further note that the DLC8 dimer represents a novel mode of three-dimensional domain swapping.  相似文献   

3.
Barbar E  Kleinman B  Imhoff D  Li M  Hays TS  Hare M 《Biochemistry》2001,40(6):1596-1605
Cytoplasmic dynein is a multisubunit ATPase that transforms chemical energy into motion along microtubules. LC8, a 10 kDa light chain subunit of the dynein complex, is highly conserved with 94% sequence identity between Drosophila and human. The precise function of this protein is unknown, but its ubiquitous expression and conservation suggest a critical role in the function of the dynein motor complex. We have overexpressed LC8 from Drosophila melanogaster and characterized its dimerization and folding using analytical ultracentrifugation, size-exclusion chromatography, circular dichroism, and fluorescence spectroscopy. Sedimentation equilibrium measurements of LC8 at pH 7 reveal a reversible monomer-dimer equilibrium with a dissociation constant of 12 microM at 4 degrees C. At lower pH, LC8 dissociates to a monomer, with a transition midpoint at pH 4.8. Far-UV CD and fluorescence spectra demonstrate that pH-dissociated LC8 retains native secondary and tertiary structures, while the diminished near-UV CD signal shows loss of quaternary structure. The observation that dimeric LC8 dissociates at low pH can be explained by titration of a histidine pair in the dimer interface. Equilibrium denaturation experiments with a protein concentration range spanning almost 2 orders of magnitude indicate that unfolding of LC8 dimer is a two-stage process, in which global unfolding is preceded by dissociation to a folded monomer. The nativelike tertiary structure of the monomer suggests a role for the monomer-dimer equilibrium of LC8 in dynein function.  相似文献   

4.
DPP-IV is a prolyl dipeptidase, cleaving the peptide bond after the penultimate proline residue. It is an important drug target for the treatment of type II diabetes. DPP-IV is active as a dimer, and monomeric DPP-IV has been speculated to be inactive. In this study, we have identified the C-terminal loop of DPP-IV, highly conserved among prolyl dipeptidases, as essential for dimer formation and optimal catalysis. The conserved residue His750 on the loop contributes significantly for dimer stability. We have determined the quaternary structures of the wild type, H750A, and H750E mutant enzymes by several independent methods including chemical cross-linking, gel electrophoresis, size exclusion chromatography, and analytical ultracentrifugation. Wild-type DPP-IV exists as dimers both in the intact cell and in vitro after purification from human semen or insect cells. The H750A mutation results in a mixture of DPP-IV dimer and monomer. H750A dimer has the same kinetic constants as those of the wild type, whereas the H750A monomer has a 60-fold decrease in kcat. Replacement of His750 with a negatively charged Glu (H750E) results in nearly exclusive monomers with a 300-fold decrease in catalytic activity. Interestingly, there is no dynamic equilibrium between the dimer and the monomer for all forms of DPP-IVs studied here. This is the first study of the function of the C-terminal loop as well as monomeric mutant DPP-IVs with respect to their enzymatic activities. The study has important implications for the discovery of drugs targeted to the dimer interface.  相似文献   

5.
Chemokine CXCL8 and its receptor CXCR1 are key mediators in combating infection and have also been implicated in the pathophysiology of various diseases including chronic obstructive pulmonary disease (COPD) and cancer. CXCL8 exists as monomers and dimers but monomer alone binds CXCR1 with high affinity. CXCL8 function involves binding two distinct CXCR1 sites – the N‐terminal domain (Site‐I) and the extracellular/transmembrane domain (Site‐II). Therefore, higher monomer affinity could be due to stronger binding at Site‐I or Site‐II or both. We have now characterized the binding of a human CXCR1 N‐terminal domain peptide (hCXCR1Ndp) to WT CXCL8 under conditions where it exists as both monomers and dimers. We show that the WT monomer binds the CXCR1 N‐domain with much higher affinity and that binding is coupled to dimer dissociation. We also characterized the binding of two CXCL8 monomer variants and a trapped dimer to two different hCXCR1Ndp constructs, and observe that the monomer binds with ~10‐ to 100‐fold higher affinity than the dimer. Our studies also show that the binding constants of monomer and dimer to the receptor peptides, and the dimer dissociation constant, can vary significantly as a function of pH and buffer, and so the ability to observe WT monomer peaks is critically dependent on NMR experimental conditions. We conclude that the monomer is the high affinity CXCR1 agonist, that Site‐I interactions play a dominant role in determining monomer vs. dimer affinity, and that the dimer plays an indirect role in regulating monomer function.  相似文献   

6.
The structure of the protein known both as neuronal nitric oxide synthase inhibitory protein, PIN (protein inhibitor of nNOS), and also as the 8 kDa dynein light chain (LC8) has been solved by X-ray diffraction. Two PIN/LC8 monomers related by a two-fold axis form a rectangular dimer. Two pairs of alpha-helices cover opposite faces, and each pair of helices packs against a beta-sheet with five antiparallel beta-strands. Each five-stranded beta-sheet contains four strands from one monomer and a fifth strand from the other monomer. A 13-residue peptide from nNOS is bound to the dimer in a deep hydrophobic groove as a sixth antiparallel beta-strand. The structure provides key insights into dimerization of and peptide binding by the multifunctional PIN/LC8 protein.  相似文献   

7.
8.
Alkaline phosphatases (APs), E.C. 3.1.3.1, are non-specific phosphomonoesterases optimally active under alkaline conditions. They are classically known to be homodimeric metalloenzymes. This quaternary structure has been considered necessary for activity, although the relationship between quaternary structure and activity is not well understood. Recombinant Pyrococcus abyssi AP was previously isolated and characterized, appearing to have two active quaternary structures on native polyacrylamide gel electrophoresis: a monomer and a homodimer. The purpose of the present work was to determine the actual quaternary structure of P. abyssi AP in solution, by isolating each of the two quaternary forms and establishing the parameters governing the assembly and dissociation of the dimer. pH appeared to be an important parameter: in acidic media, the monomer/dimer ratio shifted towards monomer. Buffer composition also affected the quaternary structure: at the same pH, in potassium phosphate buffer, the two quaternary structures were observed, whereas in tris(hydroxymethyl)aminomethane hydrochloride buffer, only the dimer was observed. Metals bound to the enzyme were found to be involved in the stability of the quaternary structure. Indeed, the P. abyssi AP obtained upon removal of the metals was monomeric. Reactivation of the latter was achieved with variable efficiency. From these experiments, no active monomer could be isolated, leading the conclusion that the active form of P. abyssi AP is the homodimer.  相似文献   

9.
Yeast (Saccharomyces cerevisiae) pyrophosphatase (Y-PPase) is a tight homodimer with two active sites separated in space from the subunit interface. The present study addresses the effects of mutation of four amino acid residues at the subunit interface on dimer stability and catalytic activity. The W52S variant of Y-PPase is monomeric up to an enzyme concentration of 300 microm, whereas R51S, H87T, and W279S variants produce monomer only in dilute solutions at pH > or = 8.5, as revealed by sedimentation, gel electrophoresis, and activity measurements. Monomeric Y-PPase is considerably more sensitive to the SH reagents N-ethylmaleimide and p-hydroxymercurobenzosulfonate than the dimeric protein. Additionally, replacement of a single cysteine residue (Cys(83)), which is not part of the subunit interface or active site, with Ser resulted in insensitivity of the monomer to SH reagents and stabilization against spontaneous inactivation during storage. Active site ligands (Mg(2+) cofactor, P(i) product, and the PP(i) analog imidodiphosphate) stabilized the W279S dimer versus monomer predominantly by decreasing the rate of dimer to monomer conversion. The monomeric protein exhibited a markedly increased (5-9-fold) Michaelis constant, whereas k(cat) remained virtually unchanged, compared with dimer. These results indicate that dimerization of Y-PPase improves its substrate binding performance and, conversely, that active site adjustment through cofactor, product, or substrate binding strengthens intersubunit interactions. Both effects appear to be mediated by a conformational change involving the C-terminal segment that generally shields the Cys(83) residue in the dimer.  相似文献   

10.
In mesophiles, triosephosphate isomerase (TIM) is an obligated homodimer. We have previously shown that monomeric folding intermediates are common in the chemical unfolding of TIM, where dissociation provides 75% of the overall conformational stability of the dimer. However, analysis of the crystallographic structure shows that, during unfolding, intermonomeric contacts contribute to only 5% of the overall increase in accessible surface area. In this work several methodologies were used to characterize the thermal dissociation and unfolding of the TIM from Entamoeba histolytica (EhTIM) and a monomeric variant obtained by chemical derivatization (mEhTIM). During EhTIM unfolding, sequential transitions corresponding to dimer dissociation into a compact monomeric intermediate followed by unfolding and further aggregation of the intermediate occurred. In the case of mEhTIM, a single transition, analogous to the second transition of EhTIM, was observed. Calorimetric, spectroscopic, hydrodynamic, and functional evidence shows that dimer dissociation is not restricted to localized interface reorganization. Dissociation represents 55% (DeltaH(Diss) = 146.8 kcal mol(-1)) of the total enthalpy change (DeltaH(Tot) = 266 kcal mol(-1)), indicating that this process is linked to substantial unfolding. We propose that, rather than a rigid body process, subunit assembly is best represented by a fly-casting mechanism. In TIM, catalysis is restricted to the dimer; therefore, the interface can be viewed as the final nucleation motif that directs assembly, folding, and function.  相似文献   

11.
The kinetics of the dissociation reaction under acidic conditions of the dimeric pig and chicken mitochondrial malate dehydrogenases (EC 1.1.1.37) have been studied. The dissociation of the pig enzyme is completely reversible. The pK for dissociation determined by light-scattering measurements agrees within experimental error with the pK value of 5.25 measured for a tyrosine-carboxylate pair. The rate constants for the dissociation reaction and for the protonation process of this tyrosine are in close agreement. Thus, the tyrosine-carboxylate pair can be used as indicator of the dissociation reaction. The dissociation of the chicken enzyme proceeds around pH 4.5 at a much lower rate. A true equilibrium between dimer and monomers is not found, since the monomer gradually unfolds at this pH. The monomers of both enzymes, pig and chicken mitochondrial malate dehydrogenase, show the same stability towards acid. The difference in stability of the dimeric forms, therefore, must be due to an altered subunit contact area.  相似文献   

12.
One of the distinctive features of eubacterial retinal-based proton pumps, proteorhodopsins, xanthorhodopsin, and others, is hydrogen bonding of the key aspartate residue, the counterion to the retinal Schiff base, to a histidine. We describe properties of the recently found eubacterium proton pump from Exiguobacterium sibiricum (named ESR) expressed in Escherichia coli, especially features that depend on Asp-His interaction, the protonation state of the key aspartate, Asp85, and its ability to accept a proton from the Schiff base during the photocycle. Proton pumping by liposomes and E. coli cells containing ESR occurs in a broad pH range above pH 4.5. Large light-induced pH changes indicate that ESR is a potent proton pump. Replacement of His57 with methionine or asparagine strongly affects the pH-dependent properties of ESR. In the H57M mutant, a dramatic decrease in the quantum yield of chromophore fluorescence emission and a 45 nm blue shift of the absorption maximum with an increase in the pH from 5 to 8 indicate deprotonation of the counterion with a pK(a) of 6.3, which is also the pK(a) at which the M intermediate is observed in the photocycle of the protein solubilized in detergent [dodecyl maltoside (DDM)]. This is in contrast with the case for the wild-type protein, for which the same experiments show that the major fraction of Asp85 is deprotonated at pH >3 and that it protonates only at low pH, with a pK(a) of 2.3. The M intermediate in the wild-type photocycle accumulates only at high pH, with an apparent pK(a) of 9, via deprotonation of a residue interacting with Asp85, presumably His57. In liposomes reconstituted with ESR, the pK(a) values for M formation and spectral shifts are 2-3 pH units lower than in DDM. The distinctively different pH dependencies of the protonation of Asp85 and the accumulation of the M intermediate in the wild-type protein versus the H57M mutant indicate that there is strong Asp-His interaction, which substantially lowers the pK(a) of Asp85 by stabilizing its deprotonated state.  相似文献   

13.
Bothopstoxin-I (BthTX-I) is a homodimeric Lys49-PLA2 homologue from the venom of Bothrops jararacussu in which a single Trp77 residue is located at the dimer interface. Intrinsic tryptophan fluorescence emission (ITFE) quenching by iodide and acrylamide has confirmed that a dimer to monomer transition occurs on reducing the pH from 7.0 to 5.0. Both the monomer and the dimer showed an excitation wavelength-dependent increase in the fluorescence emission maximum, however the excitation curve of the dimer was blue-shifted with respect to the monomeric form. No differences in the absorption or circular dichroism spectra between pH 5.0 and 7.0 were observed, suggesting that this curve shift is due neither to altered electronic ground states nor to exciton coupling of the Trp residues. We suggest that fluorescence resonance energy homotransfer between Trp77 residues at the BthTX-I dimer interface results in excitation of an acceptor Trp population which demonstrates a red-shifted fluorescence emission.  相似文献   

14.
Among manganese superoxide dismutases, residues His30 and Tyr174 are highly conserved, forming part of the substrate access funnel in the active site. These two residues are structurally linked by a strong hydrogen bond between His30 NE2 from one subunit and Tyr174 OH from the other subunit of the dimer, forming an important element that bridges the dimer interface. Mutation of either His30 or Tyr174 in Escherichia coli MnSOD reduces the superoxide dismutase activity to 30--40% of that of the wt enzyme, which is surprising, since Y174 is quite remote from the active site metal center. The 2.2 A resolution X-ray structure of H30A-MnSOD shows that removing the Tyr174-->His30 hydrogen bond from the acceptor side results in a significant displacement of the main-chain segment containing the Y174 residue, with local rearrangement of the protein. The 1.35 A resolution structure of Y174F-MnSOD shows that disruption of the same hydrogen bond from the donor side has much greater consequences, with reorientation of F174 having a domino effect on the neighboring residues, resulting in a major rearrangement of the dimer interface and flipping of the His30 ring. Spectroscopic studies on H30A, H30N, and Y174F mutants show that (like the previously characterized Y34F mutant of E. coli MnSOD) all lack the high pH transition of the wt enzyme. This observation supports assignment of the pH sensitivity of MnSOD to coordination of hydroxide ion at high pH rather than to ionization of the phenolic group of Y34. Thus, mutations near the active site, as in the Y34F mutant, as well as at remote positions, as in Y174F, similarly affect the metal reactivity and alter the effective pK(a) for hydroxide ion binding. These results imply that hydrogen bonding of the H30 imidazole N--H group plays a key role in substrate binding and catalysis.  相似文献   

15.
Dennison C  Lawler AT 《Biochemistry》2001,40(10):3158-3166
The effect of pH on Cu(I) and Cu(II) umecyanin (UCu), a phytocyanin obtained from horseradish roots, has been studied by electronic and NMR spectroscopy and using direct electrochemical measurements. A pK(a) value of approximately 9.5-9.8 is observed for the alkaline transition in UCu(II), and this leads to a slightly altered active site structure, as indicated by the changes in the paramagnetic 1H NMR spectrum. Electrochemical studies show that the pK(a) value for this transition in UCu(I) is 9.9. The alkaline transition is caused by the deprotonation of a surface lysine residue, with Lys96 being the most likely candidate. The isotropically shifted resonances in the (1)H NMR spectrum of UCu(II) also shift upon lowering the pH (pK(a) 5.8), and this can be assigned to the protonation of the surface (noncoordinating) His65 residue. This histidine titrates in UCu(I) with a pK(a) of 6.3. The reduction potential of the protein in this range is also dependent on pH, and pK(a) values matching those from NMR, for the two oxidation states of the protein, are obtained. There is no evidence for either of the active site histidines (His44 and His90) titrating in UCu(I) in the pH range studied (down to pH 3.7). Also highlighted in these studies are the remarkable active site similarities between umecyanin and the other phytocyanins which possess an axial Gln ligand.  相似文献   

16.
LC8 functions as a dimer crucial for a variety of molecular motors and non-motor complexes. Emerging models, founded on structural studies, suggest that the LC8 dimer promotes the stability and refolding of dimeric target proteins in molecular complexes, and its interactions with selective target proteins, including dynein subunits, is regulated by LC8 phosphorylation, which is proposed to prevent LC8 dimerization. To test these hypotheses in vivo, we determine the impacts of two new LC8 mutations on the assembly and stability of defined LC8-containing complexes in Chlamydomonas flagella. The three types of dyneins and the radial spoke are disparately affected by dimeric LC8 with a C-terminal extension. The defects include the absence of specific subunits, complex instability, and reduced incorporation into the axonemal super complex. Surprisingly, a phosphomimetic LC8 mutation, which is largely monomeric in vitro, is still dimeric in vivo and does not significantly change flagellar generation and motility. The differential defects in these flagellar complexes support the structural model and indicate that modulation of target proteins by LC8 leads to the proper assembly of complexes and ultimately higher level complexes. Furthermore, the ability of flagellar complexes to incorporate the phosphomimetic LC8 protein and the modest defects observed in the phosphomimetic LC8 mutant suggest that LC8 phosphorylation is not an effective mechanism for regulating molecular complexes.  相似文献   

17.
Talbott M  Hare M  Nyarko A  Hays TS  Barbar E 《Biochemistry》2006,45(22):6793-6800
Equilibrium analyses have been performed to elucidate the role of dimerization in folding and stability of dynein light chain Tctex-1. The equilibrium unfolding transition was monitored by intrinsic fluorescence intensity, fluorescence anisotropy, and circular dichroism and was modeled as a two-state mechanism where a folded dimer dissociates to two unfolded monomers without populating thermodynamically stable monomeric or dimeric intermediates. Sedimentation equilibrium and chemical cross-linking experiments performed at increasing concentrations of denaturants show no change in the association state before the unfolding transition and are consistent with the two-state model of dissociation coupled to unfolding. A linear dependence on denaturant concentration is observed by fluorescence intensity and anisotropy before unfolding in the 0-2 M GdnCl, and 0-4 M urea concentration range. This change is not protein concentration-dependent and possibly reflects relief of quenching associated with premelting conformational disorder in the vicinity of Trp 83. The data clearly show that the dissociation-coupled unfolding mechanism of Tctex-1 is different from the three-state denaturation mechanism of its structural homologue light chain LC8. The absence of a stable monomer in Tctex-1 offers insight into its functional differences from LC8.  相似文献   

18.
The villin headpiece (HP67) is a 67 residue, monomeric protein derived from the C-terminal domain of villin. Wild-type HP67 (WT HP67) is the smallest fragment of villin that retains strong in vitro actin-binding activity. WT HP67 is made up of two subdomains, which form a tightly packed interface. The C-terminal subdomain of WT HP67, denoted HP35, is rich in helical structure, folds in isolation, and has been widely used as a model system for folding studies. In contrast, very little is known about the folding of the intact villin headpiece domain. Here, NMR, CD and H/2H amide exchange measurements are used to follow the pH, thermal and urea-induced unfolding of WT HP67 and a mutant (HP67 H41Y) in which a buried conserved histidine in the N-terminal subdomain, His41, has been mutated to Tyr. Although most small proteins display two-state equilibrium unfolding, the results presented here demonstrate that unfolding of the villin headpiece is a multistate process. The presence of a folded N-terminal subdomain is shown to stabilize the C-terminal subdomain, increasing the midpoints of the thermal and urea-induced unfolding transitions and increasing protection factors for H/2H exchange. Histidine 41 has been shown to act as a pH-dependent switch in wild-type HP67: the N-terminal subdomain is unfolded when His41 is protonated, while the C-terminal subdomain remains folded irrespective of the protonation state of His41. Mutation of His41 to Tyr eliminates the segmental pH-dependent unfolding of the headpiece. The mutation stabilizes both domains, but folding is still multistate, indicating that His41 is not solely responsible for the unusual equilibrium unfolding behavior of villin headpiece domain.  相似文献   

19.
Stefin A folds as a monomer under strongly native conditions. We have observed that under partially denaturing conditions in the temperature range from 74 to 93 degrees C it folds into a dimer, while it is monomeric above the melting temperature of 95 degrees C. Below 74 degrees C the dimer is trapped and it does not dissociate. The dimer is a folded and structured protein as judged by CD and NMR, nevertheless it is no more functional as an inhibitor of cysteine proteases. The monomer-dimer transition proceeds at a slow rate and the activation energy of dimerization at 99 kcal/mol is comparable to the unfolding enthalpy. A large and negative dimerization enthalpy of -111(+/- 8) kcal/mol was calculated from the temperature dependence of the dissociation constant. An irreversible pretransition at 10-15 deg. below the global unfolding temperature has been observed previously by DSC and can now be assigned to the monomer-dimer transition. Backbone resonances of all the dimer residues were assigned using 15N isotopically enriched protein. The dimer is symmetric and the chemical shift differences between the monomer and dimer are localized around the tripartite hydrophobic wedge, which otherwise interacts with cysteine proteases. Hydrogen exchange protection factors of the residues affected by dimer formation are higher in the dimer than in the monomer. The monomer to dimer transition is accompanied by a rapid exchange of all of the amide protons which are protected in the dimer, indicating that the transition state is unfolded to a large extent. Our results demonstrate that the native monomeric state of stefin A is actually metastable but is favored by the kinetics of folding. The substantial energy barrier which separates the monomer from the more stable dimer traps each state under native conditions.  相似文献   

20.
The pH- and temperature-dependent changes in the coordination sphere of the heme c group of N alpha-acetyl microperoxidase-8 (Ac-MP-8) have been studied by examining its optical, resonance Raman, electron paramagnetic resonance, and magnetic circular dichroism spectra. An optical titration indicates that Ac-MP-8 exists in three major ionization forms over the pH 1-12 range that are linked by pK alpha values of approximately 3 and 9. The acid form that is present at pH 1.5 exists as a mixture of five- and six-coordinate high-spin species and most likely has water or buffer ions as axial ligand(s). On titration to pH 7, the His18 residue is deprotonated and becomes the proximal ligand to the iron to give a six-coordinate neutral form that has water as the sixth ligand. This form exists in a thermal high-spin intermediate-spin state equilibrium. On raising the pH to 10, an alkaline form is generated which is predominantly a five-coordinate high-spin species. It is formed by ionization of the proximal His18 residue to its imidazolate form with concomitant dissociation of the water ligand at the sixth site. At concentrations of Ac-MP-8 greater than 10 microM, some six-coordinate low-spin species are formed that are attributed to a dimer in which a His18 residue from a second molecule of Ac-MP-8 coordinates to the sixth site of another to give a bis-His complex. Raising the pH to 11.5 does not produce an appreciable amount of the six-coordinate complex with hydroxide as the sixth ligand. These studies show that Ac-MP-8 is a good water-soluble model for the peroxidases that exhibits minimal aggregation at concentrations below 10 microM in the neutral and alkaline pH regions.  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