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.
2.
A 12-residue peptide designed to form an alpha-helix and self-associate into an antiparallel 4-alpha-helical bundle yields a 0.9 A crystal structure revealing unanticipated features. The structure was determined by direct phasing with the "Shake-and-Bake" program, and contains four crystallographically distinct 12-mer peptide molecules plus solvent for a total of 479 atoms. The crystal is formed from nearly ideal alpha-helices hydrogen bonded head-to-tail into columns, which in turn pack side-by-side into sheets spanning the width of the crystal. Within each sheet, the alpha-helices run antiparallel and are closely spaced (9-10 A center-to-center). The sheets are more loosely packed against each other (13-14 A between helix centers). Each sheet is amphiphilic: apolar leucine side chains project from one face, charged lysine and glutamate side chains from the other face. The sheets are stacked with two polar faces opposing and two apolar faces opposing. The result is a periodic biomaterial composed of packed protein bilayers, with alternating polar and apolar interfaces. All of the 30 water molecules in the unit cell lie in the polar interface or between the stacked termini of helices. A section through the sheet reveals that the helices packed at the apolar interface resemble the four-alpha-helical bundle of the design, but the helices overhang parts of the adjacent bundles, and the helix crossing angles are less steep than intended (7-11 degrees rather than 18 degrees).  相似文献   

3.
Amphiphilic signal sequences with the potential to form alpha-helices with a polar, charged face and an apolar face are common in proteins which are imported into mitochondria, in the PTS permeases of bacteria, and in bacterial rhodopsins. Synthetic peptides of such sequences partition into the surface region of lipid membranes where they can adopt different secondary structures. A finely controlled balance of electrostatic and hydrophobic interactions determines the 'affinity' of amphiphilic signal peptides for lipid membranes, as well as the structure, orientation and depth of penetration of these peptides in lipid bilayer membranes. The ability of an individual peptide to associate with lipid bilayer membranes in several different modes is, most likely, a general feature of amphiphilic signal peptides and is reflected in several common physical properties of their amino acid sequences.  相似文献   

4.
The interactions of the antimicrobial peptides aurein 1.2, citropin 1.1 and maculatin 1.1 with dimyristoylphosphatidylcholine (DMPC), dimyristoylphosphatidylglycerol (DMPG) and dimyristoylphosphatidylethanolamine (DMPE) were studied by differential scanning calorimetry (DSC) and Fourier-transform infrared (FTIR) spectroscopy. The effects of these peptides on the thermotropic phase behavior of DMPC and DMPG are qualitatively similar and manifested by the suppression of the pretransition, and by peptide concentration-dependent decreases in the temperature, cooperativity and enthalpy of the gel/liquid-crystalline phase transition. However, at all peptide concentrations, anionic DMPG bilayers are more strongly perturbed than zwitterionic DMPC bilayers, consistent with membrane surface charge being an important aspect of the interactions of these peptides with phospholipids. However, at all peptide concentrations, the perturbation of the thermotropic phase behavior of zwitterionic DMPE bilayers is weak and discernable only when samples are exposed to high temperatures. FTIR spectroscopy indicates that these peptides are unstructured in aqueous solution and that they fold into α-helices when incorporated into lipid membranes. All three peptides undergo rapid and extensive H-D exchange when incorporated into D2O-hydrated phospholipid bilayers, suggesting that they are located in solvent-accessible environments, most probably in the polar/apolar interfacial regions of phospholipid bilayers. The perturbation of model lipid membranes by these peptides decreases in magnitude in the order maculatin 1.1 > aurein 1.2 > citropin 1.1, whereas the capacity to inhibit Acholeplasma laidlawii B growth decreases in the order maculatin 1.1 > aurein 1.2 ≅ citropin 1.1. The higher efficacy of maculatin 1.1 in disrupting model and biological membranes can be rationalized by its larger size and higher net charge. However, despite its smaller size and lower net charge, aurein 1.2 is more disruptive of model lipid membranes than citropin 1.1 and exhibits comparable antimicrobial activity, probably because aurein 1.2 has a higher propensity for partitioning into phospholipid membranes.  相似文献   

5.
Human arginase I is a binuclear manganese metalloenzyme that catalyzes the hydrolysis of l-arginine to generate l-ornithine and urea. We demonstrate that N-hydroxy-l-arginine (NOHA) binds to this enzyme with Kd = 3.6 μM, and nor-N-hydroxy-l-arginine (nor-NOHA) binds with Kd = 517 nM (surface plasmon resonance) or Kd ≈ 50 nM (isothermal titration calorimetry). Crystals of human arginase I complexed with NOHA and nor-NOHA afford 2.04 and 1.55 Å resolution structures, respectively, which are significantly improved in comparison with previously-determined structures of the corresponding complexes with rat arginase I. Higher resolution structures clarify the binding interactions of the inhibitors. Finally, the crystal structure of the complex with l-lysine (Kd = 13 μM) is reported at 1.90 Å resolution. This structure confirms the importance of hydrogen bond interactions with inhibitor α-carboxylate and α-amino groups as key specificity determinants of amino acid recognition in the arginase active site.  相似文献   

6.
The photophobic receptor from Natronomonas pharaonis (NpSRII) forms a photo-signalling complex with its cognate transducer (NpHtrII). In order to elucidate the complex formation in more detail, we have studied the intermolecular binding of both constituents (NpSRII and NpHtrII157; truncated at residue 157) in detergent buffers, and in lipid bilayers using FRET. The data for hetero-dimer formation of NpSRII/NpHtrII in detergent agrees well with KD values (∼ 200 nM) described in the literature. In lipid bilayers, the binding affinity between proteins in the NpSRII/NpHtrII complex is at least one order of magnitude stronger. In detergent the strength of binding is similar for both homo-dimers (NpSRII/NpSRII and NpHtrII/NpHtrII) but significantly weaker (KD  ∼ 16 μM) when compared to the hetero-dimer. The intermolecular binding is again considerably stronger in lipid bilayers; however, it is not as strong as that observed for the hetero-dimer. At a molar transducer/lipid ratio of 1:2000, which is still well above physiological concentrations, only 40% homo-dimers are formed. Apparently, in cell membranes the formation of the assumed functionally active oligomeric 2:2 complex depends on the full-length transducer including the helical cytoplasmic part, which is thought to tighten the transducer-dimer association.  相似文献   

7.
Oxidovanadium(IV), a cationic species (VO2+) of vanadium(IV), binds to several proteins, including actin. Upon titration with oxidovanadium(IV), approximately 100% quenching of the intrinsic fluorescence of monomeric actin purified from rabbit skeletal muscle (G-actin) was observed, with a V50 of 131 μM, whereas for the polymerized form of actin (F-actin) 75% of quenching was obtained and a V50 value of 320 μM. Stern-Volmer plots were used to estimate an oxidovanadium(IV)-actin dissociation constant, with Kd of 8.2 μM and 64.1 μM VOSO4, for G-actin and F-actin, respectively. These studies reveal the presence of a high affinity binding site for oxidovanadium(IV) in actin, producing local conformational changes near the tryptophans most accessible to water in the three-dimensional structure of actin. The actin conformational changes, also confirmed by 1H NMR, are accompanied by changes in G-actin hydrophobic surface, but not in F-actin. The 1H NMR spectra of G-actin treated with oxidovanadium(IV) clearly indicates changes in the resonances ascribed to methyl group and aliphatic regions as well as to aromatics and peptide-bond amide region. In parallel, it was verified that oxidovanadium(IV) prevents the G-actin polymerization into F-actin. In the 0-200 μM range, VOSO4 inhibits 40% of the extent of polymerization with an IC50 of 15.1 μM, whereas 500 μM VOSO4 totally suppresses actin polymerization. The data strongly suggest that oxidovanadium(IV) binds to actin at specific binding sites preventing actin polymerization. By affecting actin structure and function, oxidovanadium(IV) might be responsible for many cellular effects described for vanadium.  相似文献   

8.

Background

In Gram-negative bacteria the ZnuABC transporter ensures adequate zinc import in Zn(II)-poor environments, like those encountered by pathogens within the infected host. Recently, the metal-binding protein ZinT was suggested to operate as an accessory component of ZnuABC in periplasmic zinc recruitment. Since ZinT is known to form a ZinT–ZnuA complex in the presence of Zn(II) it was proposed to transfer Zn(II) to ZnuA. The present work was undertaken to test this claim.

Methods

ZinT and its structural relationship with ZnuA have been characterized by multiple biophysical techniques (X-ray crystallography, SAXS, analytical ultracentrifugation, fluorescence spectroscopy).

Results

The metal-free and metal-bound crystal structures of Salmonella enterica ZinT show one Zn(II) binding site and limited structural changes upon metal removal. Spectroscopic titrations with Zn(II) yield a KD value of 22 ± 2 nM for ZinT, while those with ZnuA point to one high affinity (KD < 20 nM) and one low affinity Zn(II) binding site (KD in the micromolar range). Sedimentation velocity experiments established that Zn(II)-bound ZinT interacts with ZnuA, whereas apo-ZinT does not. The model of the ZinT–ZnuA complex derived from small angle X-ray scattering experiments points to a disposition that favors metal transfer as the metal binding cavities of the two proteins face each other.

Conclusions

ZinT acts as a Zn(II)-buffering protein that delivers Zn(II) to ZnuA.

General significance

Knowledge of the ZinT–ZnuA relationship is crucial for understanding bacterial Zn(II) uptake.  相似文献   

9.
We report the effects of ligand presentation on the binding of aqueous proteins to solid supported lipid bilayers. Specifically, we show that the equilibrium dissociation constant can be strongly affected by ligand lipophilicity and linker length/structure. The apparent equilibrium dissociation constants (KD) were compared for two model systems, biotin/anti-biotin and 2,4-dinitrophenyl (DNP)/anti-DNP, in bulk solution and at model membrane surfaces. The binding constants in solution were obtained from fluorescence anisotropy measurements. The surface binding constants were determined by microfluidic techniques in conjunction with total internal reflection fluorescence microscopy. The results showed that the bulk solution equilibrium dissociation constants for anti-biotin and anti-DNP were almost identical, KD(bulk) = 1.7 ± 0.2 nM vs. 2.9 ± 0.1 nM. By contrast, the dissociation constant for anti-biotin antibody was three orders of magnitude tighter than for anti-DNP at a lipid membrane interface, KD = 3.6 ± 1.1 nM vs. 2.0 ± 0.2 μM. We postulate that the pronounced difference in surface binding constants for these two similar antibodies is due to differences in the ligands’ relative lipophilicity, i.e., the more hydrophobic DNP molecules had a stronger interaction with the lipid bilayers, rendering them less available to incoming anti-DNP antibodies compared with the biotin/anti-biotin system. However, when membrane-bound biotin ligands were well screened by a poly(ethylene glycol) (PEG) polymer brush, the KD value for the anti-biotin antibody could also be weakened by three orders of magnitude, 2.4 ± 1.1 μM. On the other hand, the dissociation constant for anti-DNP antibodies at a lipid interface could be significantly enhanced when DNP haptens were tethered to the end of very long hydrophilic PEG lipopolymers (KD = 21 ± 10 nM) rather than presented on short lipid-conjugated tethers. These results demonstrate that ligand presentation strongly influences protein interactions with membrane-bound ligands.  相似文献   

10.
pH-dependent (pH 6.0–8.0) quaternary structural changes of ferric Vitreoscilla hemoglobin (VHb) have been investigated using dynamic light scattering. The VHb exhibits a monomeric state under neutral conditions at pH 7.0, while the protein forms distinct homodimeric species at pH 6.0 and 8.0, respectively. The dissociation constant obtained using the Bio-Layer Interferometry technology indicates that, at pH 7.0, the monomer–monomer dissociation of VHb is about 6-fold or 5-fold higher (KD = 6.34 μM) compared with that at slightly acidic pH (KD = 1.05 μM) or slightly alkaline pH (KD = 1.22 μM). The pH-dependent absorption spectra demonstrate that the heme microenvironment of VHb is sensitive to the changes of pH value. The maximum absorption band of heme group of VHb shifts from 402 nm to 407 nm when pH changes from 6.0 to 8.0. In addition, the fluorescence emission spectra of VHb, taken at excitation wavelength of 295 nm, suggest that the single Trp122 fluorescence quantum yields in VHb are decreased due to the formation of the homodimeric species. However, the circular dichroism spectra data display that the secondary structures of VHb are little affected by pH transitions. The pH-dependent peroxidase activity of VHb was also investigated in this study. The optimum pH for VHb using 2,2′-azino-bis-(3-ethylbenzothiazoline-6-sulfonic acid) as substrate is 7.0, which implies that the monomer state of VHb would exhibit better peroxidase activity than the homodimeric species of VHb at pH 6.0 and 8.0.  相似文献   

11.
Plasmodium falciparum parasites express and traffick numerous proteins into the red blood cell (RBC), where some associate specifically with the membrane skeleton. Importantly, these interactions underlie the major alterations to the modified structural and functional properties of the parasite-infected RBC. P. falciparum Erythrocyte Membrane Protein 3 (PfEMP3) is one such parasite protein that is found in association with the membrane skeleton. Using recombinant PfEMP3 proteins in vitro, we have identified the region of PfEMP3 that binds to the RBC membrane skeleton, specifically to spectrin and actin. Kinetic studies revealed that residues 38-97 of PfEMP3 bound to purified spectrin with moderately high affinity (KD(kin) = 8.5 × 10− 8 M). Subsequent deletion mapping analysis further defined the binding domain to a 14-residue sequence (IFEIRLKRSLAQVL; KD(kin) = 3.8 × 10− 7 M). Interestingly, this same domain also bound to F-actin in a specific and saturable manner. These interactions are of physiological relevance as evidenced by the binding of this region to the membrane skeleton of inside-out RBCs and when introduced into resealed RBCs. Identification of a 14-residue region of PfEMP3 that binds to both spectrin and actin provides insight into the potential function of PfEMP3 in P. falciparum-infected RBCs.  相似文献   

12.
The complete assignments of all the proton magnetic resonance signals from each NH-CalphaH-CbetaH2 moiety in a complex peptide containing several residues of the same type has not yet been achieved without specific or stereospecific isotopic enrichment. We report the sequencing and proton magnetic resonance spectral assignments, including those of 4 aromatic residues, of tyrocidine A, an analog of the decapeptide gramicidin S. Two complementary methods, proton-proton nuclear Overhauser enhancements and scalar decoupling, evaluated by two distinct forms of difference double resonance, were used. All chemical shifts, scalar coupling constants, and [1H:1H] nuclear Overhauser enhancements for the backbone protons are reported. The [1H:1H] nuclear Overhauser enhancements are consistent with tyrocidine A possessing a beta-I turn/beta-II' turn/antiparallel beta-pleated sheet conformation. In addition to the previously proposed nuclear Overhauser enhancement criteria for beta turns and antiparallel beta sheets, another criterion for identifying the antiparallel beta sheet is demonstrated; namely, the nuclear Overhauser enhancement between 2 CalphaH protons of the central resisdues, in this case the Phe7CalphaH and Orn2CalphaH.  相似文献   

13.
Sifuvirtide, a 36 amino acid negatively charged peptide, is a novel HIV-1 fusion inhibitor with improved antiretroviral activity. In this work we evaluated the physical chemistry foundation of the interaction of sifuvirtide with biomembrane model systems. Since this peptide has aromatic residues, fluorescence spectroscopy techniques were mostly used. The interaction was assessed by partition and quenching experiments. Results showed no significant interaction with large unilamellar vesicles composed by sphingomyelin and ceramide. In contrast, sifuvirtide presented selectivity towards vesicles composed by phosphatidylcholines (PC) in the gel phase, in opposition to fluid phase PC vesicles. The interaction of this peptide with gel phase PC membranes (Kp = 1.2 × 102) is dependent on the ionic strength, which indicates the mediation of electrostatic interactions at an interfacial level. The effects of sifuvirtide on the lipid membranes' structural properties were further evaluated using dipole-potential membrane probes, zeta-potential, dynamic light scattering and atomic force microscopy measurements. The results show that sifuvirtide does not cause a noticeable effect on lipid bilayer structure, except for membranes composed by cationic phospholipids. Altogether, we can conclude that sifuvirtide presents a specific affinity towards rigid PC membranes, and the interaction is mediated by electrostatic factors, not affecting the membrane architecture.  相似文献   

14.
Irreversible inactivation of α-thrombin (T) by the serpin, heparin cofactor II (HCII), is accelerated by ternary complex formation with the glycosaminoglycans (GAGs) heparin and dermatan sulfate (DS). Low expression of human HCII in Escherichia coli was optimized by silent mutation of 27 rare codons and five secondary Shine-Dalgarno sequences in the cDNA. The inhibitory activities of recombinant HCII, and native and deglycosylated plasma HCII, and their affinities for heparin and DS were compared. Recombinant and deglycosylated HCII bound heparin with dissociation constants (KD) of 6 ± 1 and 7 ± 1 μM, respectively, ∼6-fold tighter than plasma HCII, with KD 40 ± 4 μM. Binding of recombinant and deglycosylated HCII to DS, both with KD 4 ± 1 μM, was ∼4-fold tighter than for plasma HCII, with KD 15 ± 4 μM. Recombinant HCII, lacking N-glycosylation and tyrosine sulfation, inactivated α-thrombin with a 1:1 stoichiometry, similar to plasma HCII. Second-order rate constants for thrombin inactivation by recombinant and deglycosylated HCII were comparable, at optimal GAG concentrations that were lower than those for plasma HCII, consistent with its weaker GAG binding. This weaker binding may be attributed to interference of the Asn169N-glycan with the HCII heparin-binding site.  相似文献   

15.
Bacillus brevis (ATCC 8185) produces an antibiotic peptide, tyrocidine. We found that adenosine or 5'-AMP suppressed the production of tyrocidine with half-maximum inhibition at 100-300 microM. This inhibition was specific to the production of tyrocidine since neither adenosine nor 5'-AMP showed any effect on bacterial growth. Cyclic nucleotides had no effect. These results suggest that adenosine, 5'-AMP or its metabolite was specifically involved in the regulation of tyrocidine production.  相似文献   

16.
A new radioiodinated photoaffinity compound, [125I]YE(Bpa)WSLAAPQRFNH2, derived from a peptide present in the rat neuropeptide FF (NPFF) precursor was synthesized, and its binding characteristics were investigated on a neuroblastoma clone, SH-SY5Y, stably expressing rat NPFF2 receptors tagged with the T7 epitope. The binding of the probe was saturable and revealed a high-affinity interaction (KD = 0.24 nM) with a single class of binding sites. It was also able to affinity label NPFF2 receptor in a specific and efficient manner given that 38% of the bound radioligand at saturating concentration formed a wash-resistant binding after ultraviolet (UV) irradiation. Photoaffinity labeling with [125I]YE(Bpa)WSLAAPQRFamide showed two molecular forms of NPFF2 receptor with apparent molecular weights of 140 and 95 kDa in a 2:1 ratio. The comparison of the results between photoaffinity labeling and Western blot analysis suggests that all receptor forms bind the probe irreversibly with the same efficiency. On membranes of mouse olfactory bulb, only the high molecular weight form of NPFF2 receptor is observed. [125I]YE(Bpa)WSLAAPQRFamide is an excellent radioiodinated peptidic ligand for direct and selective labeling of NPFF2 receptors in vitro.  相似文献   

17.
The interactions between Ca2+ and C-reactive protein (CRP) have been characterized using a surface plasmon resonance (SPR) biosensor. The protein was immobilized on a sensor chip, and increasing concentrations of Ca2+ or phosphocholine were injected. Binding of Ca2+ induced a 10-fold higher signal than expected from the molecular weight of Ca2+. It was interpreted to result from the conformational change that occurs on binding of Ca2+. Two sites with different characteristics were distinguished: a high-affinity site with KD = 0.03 mM and a low-affinity site with KD = 5.45 mM. The pH dependencies of the two Ca2+ interactions were different and enabled the assignment of the different sites in the three-dimensional structure of CRP. There was no evidence for cooperativity in the phosphocholine interaction, which had KD = 5 μM at 10 mM Ca2+. SPR biosensors can clearly detect and quantify the binding of very small molecules or ions to immobilized proteins despite the theoretically very low signals expected on binding, provided that significant conformational changes are involved. Both the interactions and the conformational changes can be characterized. The data have important implications for the understanding of the function of CRP and suggest that Ca2+ is an efficient regulator under physiological conditions.  相似文献   

18.
The cationic β-sheet cyclic tetradecapeptide cyclo[VKLdKVdYPLKVKLdYP] (GS14dK4) is a diastereomeric lysine ring-size analog of the potent naturally occurring antimicrobial peptide gramicidin S (GS) which exhibits enhanced antimicrobial but markedly reduced hemolytic activity compared to GS itself. We have previously studied the binding of GS14dK4 to various phospholipid bilayer model membranes using isothermal titration calorimetry [Abraham, T. et al. (2005) Biochemistry 44, 2103-2112]. In the present study, we compare the ability of GS14dK4 to bind to and disrupt these same phospholipid model membranes by employing a fluorescent dye leakage assay to determine the ability of this peptide to permeabilize large unilamellar vesicles. We find that in general, the ability of GS14dK4 to bind to and to permeabilize phospholipid bilayers of different compositions are not well correlated. In particular, the binding affinity of GS14dK4 varies markedly with the charge and to some extent with the polar headgroup structure of the phospholipid and with the cholesterol content of the model membrane. Specifically, this peptide binds much more tightly to anionic than to zwitterionic phospholipids and much less tightly to cholesterol-containing than to cholesterol-free model membranes. In addition, the maximum extent of binding of GS14dK4 can also vary considerably with phospholipid composition in a parallel fashion. In contrast, the ability of this peptide to permeabilize phospholipid vesicles is only weakly dependent on phospholipid charge, polar headgroup structure or cholesterol content. We provide tentative explanations for the observed lack of a correlation between the affinity and extent of GS14dK4 binding to, and degree of disruption of the structure and integrity of, phospholipid bilayers membranes. We also present evidence that the lack of correlation between these two parameters may be a general phenomenon among antimicrobial peptides. Finally, we demonstrate that the affinity of binding of GS14dK4 to various phospholipid bilayer membranes is much more strongly correlated with the antimicrobial and hemolytic activities of this peptide than with its effect on the rate and extent of dye leakage in these model membrane systems.  相似文献   

19.
The impact of Zn2+ ion on interactions of flavonols galangin (Gal), kaempferol (Kae), quercetin (Que) and myricetin (Myr) with bovine serum albumin (BSA) in aqueous solution were studied by fluorescence quenching technique. The results exhibited that Zn2+ ion affected significantly the interactions and the effect was distinct for the flavonol bearing different number of B-ring hydroxyl. Each flavonol can quench the fluorescence of BSA, displaying a quenching extent of Myr > Que > Kae > Gal, which is in good agreement with the number variation of the B-ring hydroxyl. The presence of Zn2+ ion promoted the quenching for the flavonols, exhibiting an extent of Que > Myr > Kae > Gal. The values of Ka for Kae, Que and Myr decreased whereas KSV and kq for Gal, Kae and Que increased with the number of B-ring hydroxyl. The type of BSA fluorescence quenching for Gal, Kae and Que hardly changed but the preference of static quenching increased. The values of KSV and kq for Myr remarkably decreased and the fluorescence quenching of BSA alternatively occurred via both static and dynamic type instead of only one (static or dynamic). The results suggest the key role of the B-ring hydroxyl and the distinct effect of its number in the interactions. Each flavonol may capture the BSA-bound ZnII in the solution, forming ZnII-flavonol complex that is possibly responsible for BSA fluorescence quenching. The B-ring hydroxyl could establish hydrogen bonds with BSA in the absence of Zn2+ and act as donors for chelating in the presence of Zn2+. The formation of dinuclear ZnII-Myr complex together with the hydrogen bonds between the free B-ring hydroxyl and BSA may contribute to the exceptional behavior of Myr.  相似文献   

20.
Summary The influence of fructose 2,6-bisphosphate on the activation of purified swine kidney phosphofructokinase as a function of the concentration of fructose 6P, ATP and citrate was investigated. The purified enzyme was nearly completely inhibited in the presence of 2 mM ATP. The addition of 20 nM fructose 2,6-P2 reversed the inhibition and restored more than 80% of the activity. In the absence of fructose 2,6-P2 the reaction showed a sigmoidal dependence on fructose 6-phosphate. The addition of 10 nM fructose 2,6-bisphosphate decreased the K0.5 for fructose 6-phosphate from 3 mM to 0.4 mM in the presence of 1.5 mM ATP. These results clearly show that fructose 2,6-bisphosphate increases the affinity of the enzyme for fructose 6-phosphate and decreases the inhibitory effect of ATP. The extent of inhibition by citrate was also significantly decreased in the presence of fructose 2,6-phosphate.The influence of various effectors of phosphofructokinase on the binding of ATP and fructose 6-P to the enzyme was examined in gel filtration studies. It was found that kidney phosphofructokinase binds 5.6 moles of fructose 6-P per mole of enzyme, which corresponds to about one site per subunit of tetrameric enzyme. The KD for fructose 6-P was 13 µM and in the presence of 0.5 mM ATP it increased to 27 µM. The addition of 0.3 mM citrate also increased the KD for fructose 6-P to about 40 µM. AMP, 10 µM, decreased the KD to 5 µM and the addition of fructose 2,6-phosphate decreased the KD for fructose 6-P to 0.9 µM. The addition of these compounds did not effect the maximal amount of fructose 6-P bound to the enzyme, which indicated that the binding site for these compounds might be near, but was not identical to the fructose 6-P binding site. The enzyme bound a maximum of about 12.5 moles of ATP per mole, which corresponds to 3 moles per subunit. The KD of the site with the highest affinity for ATP was 4 µM, and it increased to 15 µM in the presence of fructose 2,6-bisphosphate. The addition of 50 µM fructose 1,6-bisphosphate increased the KD for ATP to 5.9 µM. AMP increased the KD to 5.9 µM whereas 0.3 mM citrate decreased the KD for ATP to about 2 µM. The KD for AMP, was 2.0 µM; the KD for cyclic AMP was 1.0 µM; the KD for ADP was 0.9 µM; the KD for fructose 1,6-bisphosphate was 0.5 µM; the KD for citrate was 0.4 µM and the KD for fructose 2,6-bisphosphate was about 0.1 µM. A maximum of about 4 moles of AMP, ADP and cyclic AMP and fructose 2,6-bisphosphate were bound per mole of enzyme. Taken collectively, these and previous studies (9) indicate that fructose 2,6-phosphate is a very effective activator of swine kidney phosphofructokinase. This effector binds to the enzyme with a very high affinity, and significantly decreases the binding of ATP at the inhibitory site on the enzyme.  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