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 pH-dependent kinetics of lysyl oxidase catalysis was examined for evidence of an ionizable enzyme residue which might function as a general base catalyzing proton abstraction previously shown to be a component of the mechanism of substrate processing by this enzyme. Plots of log Vmax/Km for the oxidation of n-hexylamine versus pH yielded pKa values of 7.0 +/- 0.1 and 10.4 +/- 0.1. The higher pKa varied with different substrates, reflecting ionization of the substrate amino group. A van't Hoff plot of the temperature dependence of the lower pKa yielded a value of 6.1 kcal mol-1 for the enthalpy of ionization. This value as well as the pKa of 7.0 are consistent with those of histidine residues previously implicated as general base catalysts in enzymes. Incubation of lysyl oxidase with low concentrations of diethyl pyrocarbonate, a histidine-selective reagent, at 22 degrees C and pH 7.0 irreversibly inhibited enzyme activity by a pseudo first-order kinetic process. The inactivation of lysyl oxidase correlated with spectral and pH-dependent kinetic evidence for the chemical modification of 1 histidine residue/mol of enzyme, the pKa of which was 6.9 +/- 0.1, within experimental error of that seen in the plot of log Vmax/Km versus pH. Enzyme activity was restored by incubation of the modified enzyme with hydroxylamine, consistent with the ability of this nucleophile to displace the carbethoxy group from N-carbethoxyhistidine. The presence of the n-hexylamine substrate largely protected against enzyme inactivation by diethyl pyrocarbonate. These results thus indicate a functional role for histidine in lysyl oxidase catalysis consistent with that of a general base in proton abstraction.  相似文献   

2.
Bose K  Pop C  Feeney B  Clark AC 《Biochemistry》2003,42(42):12298-12310
We have examined the enzymatic activity of an uncleavable procaspase-3 mutant (D9A/D28A/D175A), which contains the wild-type catalytic residues in the active site. The results are compared to those for the mature caspase-3. Although at pH 7.5 and 25 degrees C the K(m) values are similar, the catalytic efficiency (k(cat)) is approximately 130-fold lower in the zymogen. The mature caspase-3 demonstrates a maximum activity at pH 7.4, whereas the maximum activity of procaspase-3 occurs at pH 8.3. The pK(a) values of both catalytic groups, H121 and C163, are shifted to higher pH for procaspase-3. We developed limited proteolysis assays using trypsin and V8 proteases, and we show that these assays allow the examination of amino acids in three of five active site loops. In addition, we examined the fluorescence emission of the two tryptophanyl residues in the active site over the pH range of 2.5-9 as well as the response to several quenching agents. Overall, the data suggest that the major conformational change that occurs upon maturation results in formation of the loop bundle among loops L4, L2, and L2'. The pK(a) values of both catalytic groups decrease as a result of the loop movements. However, loop L3, which comprises the bulk of the substrate binding pocket, does not appear to be unraveled and solvent-exposed, even at lower pH.  相似文献   

3.
The direct induction of apoptosis has emerged as a powerful anticancer strategy, and small molecules that either inhibit or activate certain proteins in the apoptotic pathway have great potential as novel chemotherapeutic agents. Central to apoptosis is the activation of the zymogen procaspase-3 to caspase-3. Caspase-3 is the key “executioner” caspase, catalyzing the hydrolysis of a multitude of protein substrates within the cell. Interestingly, procaspase-3 levels are often elevated in cancer cells, suggesting a compound that directly stimulates the activation of procaspase-3 to caspase-3 could selectively induce apoptosis in cancer cells. We recently reported the discovery of a compound, PAC-1, which enhances procaspase-3 activity in vitro and induces apoptotic death in cancer cells in culture and in mouse xenograft models. Described herein is the mechanism by which PAC-1 activates procaspase-3 in vitro. We show that zinc inhibits the enzymatic activity of procaspase-3 and that PAC-1 strongly activates procaspase-3 in buffers that contain zinc. PAC-1 and zinc form a tight complex with one another, with a dissociation constant of approximately 42 nM. The combined data indicate that PAC-1 activates procaspase-3 in vitro by sequestering inhibitory zinc ions, thus allowing procaspase-3 to autoactivate itself to caspase-3. The small-molecule-mediated activation of procaspases has great therapeutic potential and thus this discovery of the in vitro mechanism of action of PAC-1 is critical to the development and optimization of other procaspase-activating compounds.  相似文献   

4.
Sarcosine oxidase from Corynebacterium sp. U-96 is inactivated by iodoacetamide with the modification of two specific residues. Comparing the amino acid sequence and mass spectra of the peptide fragments containing the modified residues with those from the native enzyme, the modified residues were identified to be lysine. The pKa of these residues were estimated to be 8.5 and 6.7 from the pH dependence of inactivation in the presence and absence of the competitive inhibitor, acetate. These estimated pKa values are much lower than that of the epsilon-amino group of lysine residue. There may be unique microenvironments around these residues that activate their -amino groups to be susceptible to iodoacetamide. A possible role of the lysine residue with pKa 6.7 is discussed.  相似文献   

5.
Although early studies of inhibitor of apoptosis proteins (IAPs) suggested that cIAP1 directly binds and inhibits caspases similarly to X-linked IAP (XIAP), a recent one found that micromolar concentrations of cIAP1 only weakly inhibit caspase-3, -7, or -9. Here, we show that cIAP1 specifically and cooperatively blocks the cytochrome c-dependent apoptosome in vitro. Hence, cIAP1 prevented the activation of procaspase-3 but had no effect on the processing of procaspase-9 or the activity of prior activated caspase-3. Like cIAP1, XIAP had no effect on procaspase-9 processing and was a more potent inhibitor of procaspase-3 activation than of already activated caspase-3 activity. Inhibition of procaspase-3 activation depended on BIR2 and BIR3 of cIAP1 and was independent of BIR1, RING, CARD, and UBA domains. Smac prevented cIAP1 from inhibiting procaspase-3 activation and reversed the inhibition by prior addition of cIAP1. A procaspase-9 mutant (D315A) that cannot produce the p12 subunit was resistant to inhibition by cIAP1. Therefore, the N-terminal Ala-Thr-Pro-Phe motif of the p12 subunit of the caspase-9 apoptosome facilitates apoptosome blockade. Consequently, cIAP1 cooperatively interacts with oligomerized processed caspase-9 in the apoptosome and blocks procaspase-3 activation.  相似文献   

6.
The activation of caspases is a central step in apoptosis and may also be critical for terminal differentiation of epidermal keratinocytes (KC). In particular, caspase-3 has been implicated in the differentiation of embryonic KC as well as in programmed cell death of KC, and caspase-14 has been suggested to function in the formation or homeostasis of the stratum corneum (SC). To test the putative roles of these proteases, we determined their expression level and activation status during development of fetal mouse epidermis. The level of procaspase-3 did not change significantly during epidermal development, and enzyme activation was undetectable at any timepoint investigated. Despite the lack of active caspase-3, the newly formed stratum granulosum and the regressing periderm contained cells positive in the terminal deoxynucleotidyl transferase-mediated fluorescein-dUTP nick end labeling assay, indicating that nuclear DNA was degraded without activation of caspase-3, thereby arguing against a proteolytic function of caspase-3 in embryonic KC differentiation. By contrast, caspase-14 increased in abundance from embryonic day 14.5 (E14.5) onwards and consistently localized to the suprabasal layers of fetal epidermis. The caspase-14 pro-enzyme was processed into its catalytic subunits, a step required for enzyme activity, on day E17.5, coinciding with SC formation. Thus, processing of procaspase-14 is not confined to air-exposed mature skin but also occurs during epidermal development in utero. In summary, this study demonstrates that caspase-14, but not caspase-3 activation coincides temporally and spatially with embryonic KC differentiation, suggesting a role for caspase-14 in terminally differentiated KC.  相似文献   

7.
Many proteins and bioactive peptides contain an N-terminal pyroglutamate residue (Pyr1). This residue reduces the susceptibility of the protein to aminopeptidases and often has important functional roles. The antitumor ribonuclease RC-RNase 3 (RNase 3) from oocytes of Rana catesbeiana (bullfrog) is one such protein. We have produced recombinant RNase 3 containing the N-terminal Pyr1 (pRNase 3) and found it to be indistinguishable from the native RNase 3 by mass spectrometry and a variety of other biochemical and immunological criteria. We demonstrated by NMR analysis that the Pyr1 of pRNase 3 forms hydrogen bonds with Lys9 and Ile96 and stabilizes the N-terminal alpha-helix in a rigid conformation. In contrast, the N-terminal alpha-helix becomes flexible and the pKa values of the catalytic residues His10 and His97 altered when Pyr1 formation is blocked by an extra methionine at the N terminus in the recombinant mqRNase 3. Thus, our results provide a mechanistic explanation on the essential role of Pyr1 in maintaining the structural integrity, especially at the N-terminal alpha-helix, and in providing the proper environment for the ionization of His10 and His97 residues for catalysis and cytotoxicity against HeLa cells.  相似文献   

8.
Enantiomers of 4-nitrophenyl 4-X-phenacyl methylphosphonate esters (X = H, PMN; CH3 and CH3O) inactivate human factor Xa with rate constants 8-86 M(-1)s(-1) at pH 6.75 in 0.025 M Hepes buffer, 0.15 M NaCl and 2 mM CaCl2 at 7.0+/-0.1 degrees C. The stereoselectivity of the inactivation of factor Xa is 2-10 and favors the levorotatory enantiomers. The pH-dependence of inactivation of factor Xa by (-)-PMN is sigmoidal and consistent with the participation of a catalytic residue with a pKa of 6.2+/-0.1. Factor Xa reactivates from its phosphonyl adducts through a self-catalyzed intramolecular reaction, which is much influenced by the presence of phospholipids. The rate of reactivation in the absence of phospholipids is not pH dependent at pH <9, but it increases very much at pH >9. In the presence of phospholipids, the pH dependence of the rate constant for reactivation is sigmoidal in the pH 6.5-10.3 range and levels off at pH >9 indicating that the enzyme catalyzes its reactivation. The kinetic pKa for the recovery of factor Xa from its adducts with the PMNs is in the range of 6.7-8.1 and is consistent with the participation of the catalytic His57 in the reactivation process.  相似文献   

9.
Most proteases are expressed as inactive precursors, or zymogens, that become activated by limited proteolysis. We previously identified a small molecule, termed 1541, that dramatically promotes the maturation of the zymogen, procaspase-3, to its mature form, caspase-3. Surprisingly, compound 1541 self-assembles into nanofibrils, and localization of procaspase-3 to the fibrils promotes activation. Here, we interrogate the biochemical mechanism of procaspase-3 activation on 1541 fibrils in addition to proteogenic amyloid-β(1–40) fibrils. In contrast to previous reports, we find no evidence that procaspase-3 alone is capable of self-activation, consistent with its fate-determining role in executing apoptosis. In fact, mature caspase-3 is >107-fold more active than procaspase-3, making this proenzyme a remarkably inactive zymogen. However, we also show that fibril-induced colocalization of trace amounts of caspase-3 or other initiator proteases with procaspase-3 dramatically stimulates maturation of the proenzyme in vitro. Thus, similar to known cellular signaling complexes, these synthetic or natural fibrils can serve as platforms to concentrate procaspase-3 for trans-activation by upstream proteases.  相似文献   

10.
The shikimate pathway enzyme 3-dehydroquinase is very susceptible to inactivation by the group-specific reagent diethyl pyrocarbonate (DEP). Inactivation follows pseudo first-order kinetics and exhibits a second-order rate constant of 148.5 M-1 min-1. An equilibrium mixture of substrate and product substantially protects against inactivation by DEP, suggesting that residues within the active site are being modified. Complete inactivation of the enzyme correlates with the modification of 6 histidine residues/subunit as determined by difference spectroscopy at 240 nm. Enzymic activity can be restored by hydroxylamine treatment, which is also consistent with the modification occurring at histidine residues. Using the kinetic method of Tsou (Tsou, C.-L. (1962) Sci. Sin. 11, 1535-1558), it was shown that modification of a single histidine residue leads to inactivation. Ligand protection experiments also indicated that 1 histidine residue was protected from DEP modification. pH studies show that the pKa for this inactivation is 6.18, which is identical to the single pKa determined from the pH/log Vmax profile for the enzyme. A single active site peptide was identified by differential peptide mapping in the presence and absence of ligand. This peptide was found to comprise residues 141-158; of the 2 histidines in this peptide (His-143 and His-146), only one, His-143, is conserved among all type I dehydroquinases. We propose that His-143 is the active site histidine responsible for DEP-mediated inactivation of dehydroquinase and is a good candidate for the general base that has been postulated to participate in the mechanism of this enzyme.  相似文献   

11.
Chemical modification studies were performed to elucidate the role of Cys-residues in the catalysis/binding of restriction endonuclease Cfr9I. Incubation of restriction endonuclease Cfr9I with N-ethylmaleimide (NEM), iodoacetate, 5,5'-dithiobis (2-nitrobenzoic acid) at pH 7.5 led to a complete loss of the catalytic activity. However, no enzyme inactivation was detectable after modification of the enzyme with iodoacetamide and methyl methanethiosulfonate. Complete protection of the enzyme against inactivation by NEM was observed in the presence of substrate implying that Cys-residues may be located at or in the vicinity of the active site of enzyme. Direct substrate-binding studies of native and modified restriction endonuclease Cfr9I using a gel-mobility shift assay indicated that the modification of the enzyme by NEM was hindered by substrate binding. A single Cys-residue was modified during the titration of the enzyme with DTNB with concomitant loss of the catalytic activity. The pH-dependence of inactivation of Cfr9I by NEM revealed the modification of the residue with the pKa value of 8.9 +/- 0.2. The dependence of the reaction rate of substrate hydrolysis by Cfr9I versus pH revealed two essential residues with pKa values of 6.3 +/- 0.15 and 8.7 +/- 0.15, respectively. The evidence presented suggests that the restriction endonuclease Cfr9I contains a reactive sulfhydryl residue which is non-essential for catalysis, but is located at or near the substrate binding site.  相似文献   

12.
Catalytic role of histidine 147 in Escherichia coli thymidylate synthase   总被引:3,自引:0,他引:3  
Nine mutant thymidylate synthases were isolated that only differed in sequence at position 147. The wild-type enzyme (which had a histidine residue at 147) and mutant enzymes were purified to near homogeneity and their kinetic properties were compared. Although the kcat values for the mutant enzymes were 10-10,000-fold lower than for the wild-type enzyme, the Km values for both 2'-deoxyuridylate and 5,10-methylenetetrahydrofolate were nearly identical for all the enzymes indicating that His-147 is not significantly involved in initial substrate binding. By comparing the wild-type (His-147) to the glycine (Gly-147) enzyme, the side chain of His-147 was estimated to lower the activation energy of the catalytic step by 1.6-2.9 kcal mol-1. In contrast to the wild-type enzyme, the activity of the Gly-147 enzyme decreased when the pH was raised above 7.5. The activity loss coincided with the deprotonation of a residue that had a pKa of 9.46 +/- 0.2 and an enthalpy of ionization (delta Hion) of 12.1 +/- 0.9. These values are consistent with the involvement of a lysine or an arginine residue in the catalytic process. An inspection of the rates of ternary complex formation among enzyme, 5-fluoro-2'-deoxyuridylate, and 5,10-methylenetetrahydrofolate for the mutant enzymes indicated that His-147 is not needed for the proton removal from C-5 of 2'-deoxyuridylate but rather participates in an initial catalytic step and alters the pKa value of a catalytically important lysine or arginine residue.  相似文献   

13.
Drosophila alcohol dehydrogenase belongs to the short chain dehydrogenase/reductase (SDR) family which lack metal ions in their active site. In this family, it appears that the three amino acid residues, Ser138, Tyr151 and Lys155 have a similar function as the catalytic zinc in medium chain dehydrogenases. The present work has been performed in order to obtain information about the function of these residues. To obtain this goal, the pH and temperature dependence of various kinetic coefficients of the alcohol dehydrogenase from Drosophila lebanonensis was studied and three-dimensional models of the ternary enzyme-coenzyme-substrate complexes were created from the X-ray crystal coordinates of the D. lebanonensis ADH complexed with either NAD(+) or the NAD(+)-3-pentanone adduct. The kon velocity for ethanol and the ethanol competitive inhibitor pyrazole increased with pH and was regulated through the ionization of a single group in the binary enzyme-NAD(+) complex, with a DeltaHion value of 74(+/-4) kJ/mol (18(+/-1) kcal/mol). Based on this result and the constructed three-dimensional models of the enzyme, the most likely candidate for this catalytic residue is Ser138. The present kinetic study indicates that the role of Lys155 is to lower the pKa values of both Tyr151 and Ser138 already in the free enzyme. In the binary enzyme-NAD(+) complex, the positive charge of the nicotinamide ring in the coenzyme further lowers the pKa values and generates a strong base in the two negatively charged residues Ser138 and Tyr151. With the OH group of an alcohol close to the Ser138 residue, an alcoholate anion is formed in the ternary enzyme NAD(+) alcohol transition state complex. In the catalytic triad, along with their effect on Ser138, both Lys155 and Tyr151 also appear to bind and orient the oxidized coenzyme.  相似文献   

14.
3-Ketovalidoxylamine A C-N lyase of Flavobacterium saccharophilum is a monomeric protein with a molecular weight of 36,000. Amino acid analysis revealed that the enzyme contains 5 histidine residues and no cysteine residue. The enzyme was inactivated by diethylpyrocarbonate (DEP) following pseudo-first order kinetics. Upon treatment of the inactivated enzyme with hydroxylamine, the enzyme activity was completely restored. The difference absorption spectrum of the modified versus native enzyme exhibited a prominent peak around 240 nm, but there was no absorbance change above 270 nm. The pH-dependence of inactivation suggested the involvement of an amino acid residue having a pKa of 6.8. These results indicate that the inactivation is due to the modification of histidine residues. Substrates of the lyase, p-nitrophenyl-3-ketovalidamine, p-nitrophenyl-alpha-D-3-ketoglucoside, and methyl-alpha-D-3-ketoglucoside, protected the enzyme against the inactivation, suggesting that the modification occurred at or near the active site. Although several histidine residues were modified by DEP, a plot of log (reciprocal of the half-time of inactivation) versus log (concentration of DEP) suggested that one histidine residue has an essential role in catalysis.  相似文献   

15.
Coxsackievirus B3-induced apoptosis and caspase-3   总被引:11,自引:0,他引:11  
Yuan JP  Zhao W  Wang HT  Wu KY  Li T  Guo XK  Tong SQ 《Cell research》2003,13(3):203-209
  相似文献   

16.
Proteinase 3 (PR3) is an abundant serine protease of neutrophil granules and a major target of autoantibodies (PR3 anti-neutrophil cytoplasmic antibodies) in granulomatosis with polyangiitis. Some of the PR3 synthesized by promyelocytes in the bone marrow escapes the targeting to granules and occurs on the plasma membrane of naive and primed neutrophils. This membrane-associated PR3 antigen may represent pro-PR3, mature PR3, or both forms. To discriminate between mature PR3 and its inactive zymogen, which have different conformations, we generated and identified a monoclonal antibody called MCPR3-7. It bound much better to pro-PR3 than to mature PR3. This monoclonal antibody greatly reduced the catalytic activity of mature PR3 toward extended peptide substrates. Using diverse techniques and multiple recombinant PR3 variants, we characterized its binding properties and found that MCPR3-7 preferentially bound to the so-called activation domain of the zymogen and changed the conformation of mature PR3, resulting in impaired catalysis and inactivation by α1-proteinase inhibitor (α1-antitrypsin). Noncovalent as well as covalent complexation between PR3 and α1-proteinase inhibitor was delayed in the presence of MCPR3-7, but cleavage of certain thioester and paranitroanilide substrates with small residues in the P1 position was not inhibited. We conclude that MCPR3-7 reduces PR3 activity by an allosteric mechanism affecting the S1′ pocket and further prime side interactions with substrates. In addition, MCPR3-7 prevents binding of PR3 to cellular membranes. Inhibitory antibodies targeting the activation domain of PR3 could be exploited as highly selective inhibitors of PR3, scavengers, and clearers of the PR3 autoantigen in granulomatosis with polyangiitis.  相似文献   

17.
Caspase activation during apoptosis occurs in a cascade from the initiator caspase(s) (e.g. caspase-8) to the effector caspases (e.g. caspase-3), which ensures the generation of large amounts of active caspases to dismantle cells. However, the mechanism that safeguards against inadvertent caspase activation is not well understood. Previous studies have suggested that the activation of procaspase-8 is mediated by cross-cleavage of precursor dimers, formed upon apoptosis induction, which are not only enzymatically competent but also highly susceptible to cleavage, and that procaspase-8 activation is a linear process without self-amplification. Effector procaspases constitutively exist as dimers and their activation is started by trans-cleavage by an initiator caspase followed by autocleavage of effector caspases. Here we show that the dimerization of caspase-3 molecules through their protease domains is required for their processing by initiator caspases. The subsequent autoprocessing takes place through cleavage between the dimeric intermediates. Moreover, mature caspase-3 fails to process its own precursor. Thus, despite a marked difference in the generation of active intermediates, the activation of initiator and effector caspases shares the features of interdimer cleavage and lack of self-amplification. These features may be important in preventing accidental cell death.  相似文献   

18.
Treatment of 26L cells, a subclone obtained from U937 cells, with TNF-alpha or DNA-damaging agents such as teniposide (VM26) and camptothecin (CPT) induced morphologically and biochemically typical apoptotic changes, including the activation of procaspase-3. The cells persistently infected with HIV-1 (26L/HIV), however, showed a marked resistance to VM26 and CPT, whereas they hardly lost the sensitivity to TNF-alpha. TNF-alpha-induced apoptosis of 26L/HIV cells proceeded without the increase in caspase-3 activity, indicating that signaling for apoptosis in the infected cells proceeded through an alternative caspase-3-independent pathway which could respond to TNF-alpha but not to VM26 and CPT. The evidence that p-toluenesulfonyl-l-lysine chloromethyl ketone (a trypsin-like serine protease inhibitor) blocked VM26- and CPT-induced apoptotic changes but not TNF-alpha-induced apoptosis also supported the existence of the alternative TNF-alpha-inducible pathway. The results also suggest that a TLCK-sensitive protease is involved upstream of the procaspase-3 activation process and that the protease is essential for the progress of VM26- and CPT-induced apoptosis. The similar effect of HIV-1-productive infection on the apoptosis induced by the DNA-damaging agents was also confirmed by utilizing U1 cells, which are latently HIV-1-infected U937 cells. The cells became resistant to these agents after induction of the viral production by pretreatment with PMA. These results suggest that persistent HIV-1 infection blocks an apoptotic pathway triggered by DNA damaging agents through the inhibition of the procaspase-3 activation process.  相似文献   

19.
The intrinsic apoptosis apparatus plays a significant role in generating and amplifying cell death signals. In this study we examined whether there are differences in the expression of its components and in its functioning in non-small cell lung carcinoma (NSCLC) and the lung. We show that NSCLC cell lines express Apaf-1 and procaspase-9 and -3 proteins and that the expression of Apaf-1 and procaspase-3, but not of procaspase-9 and -7, is frequently up-regulated in NSCLC tissues as compared to the lung. NSCLC tissues and lungs and some NSCLC cell lines expressed also caspase-9S(b) and displayed a high caspase-9S(b)/procaspase-9 expression ratio. Procaspase-3 from NSCLCs and lungs was readily processed to caspase-3 by granzyme B or caspase-8, and the granzyme B-generated caspase-3-like activity was significantly higher in tumor tissues and cells than in lungs. By contrast, cytochrome c plus dATP could induce a significant increase of caspase-3-like activity in cytosol only in some NSCLC cell lines and in subsets of studied NSCLC tissues and lungs, while procaspase-3 and -7 were detectably processed only in NSCLC tissues which showed a high (cytochrome c+dATP)-induced caspase-3-like activity. Taken together, the present study provides evidence that the expression of Apaf-1 and procaspase-3 is up-regulated in NSCLCs and indicates that the tumors have a capability to suppress the apoptosome-driven caspase activation in their cytosol.  相似文献   

20.
Yin Q  Park HH  Chung JY  Lin SC  Lo YC  da Graca LS  Jiang X  Wu H 《Molecular cell》2006,22(2):259-268
Caspase-9 activation is critical for intrinsic cell death. The activity of caspase-9 is increased dramatically upon association with the apoptosome, and the apoptosome bound caspase-9 is the caspase-9 holoenzyme (C9Holo). In this study, we use quantitative enzymatic assays to fully characterize C9Holo and a leucine-zipper-linked dimeric caspase-9 (LZ-C9). We surprisingly show that LZ-C9 is more active than C9Holo for the optimal caspase-9 peptide substrate LEHD-AFC but is much less active than C9Holo for the physiological substrate procaspase-3. The measured Km values of C9Holo and LZ-C9 for LEHD-AFC are similar, demonstrating that dimerization is sufficient for catalytic activation of caspase-9. The lower activity of C9Holo against LEHD-AFC may be attributed to incomplete C9Holo assembly. However, the measured Km of C9Holo for procaspase-3 is much lower than that of LZ-C9. Therefore, in addition to dimerization, the apoptosome activates caspase-9 by enhancing its affinity for procaspase-3, which is important for procaspase-3 activation at the physiological concentration.  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