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 0 毫秒
1.
Mutation and aberrant expression of apoptotic proteins are hallmarks of cancer. These changes prevent proapoptotic signals from being transmitted to executioner caspases, thereby averting apoptotic death and allowing cellular proliferation. Caspase-3 is the key executioner caspase, and it exists as an inactive zymogen that is activated by upstream signals. Notably, concentrations of procaspase-3 in certain cancerous cells are significantly higher than those in noncancerous controls. Here we report the identification of a small molecule (PAC-1) that directly activates procaspase-3 to caspase-3 in vitro and induces apoptosis in cancerous cells isolated from primary colon tumors in a manner directly proportional to the concentration of procaspase-3 inside these cells. We found that PAC-1 retarded the growth of tumors in three different mouse models of cancer, including two models in which PAC-1 was administered orally. PAC-1 is the first small molecule known to directly activate procaspase-3 to caspase-3, a transformation that allows induction of apoptosis even in cells that have defective apoptotic machinery. The direct activation of executioner caspases is an anticancer strategy that may prove beneficial in treating the many cancers in which procaspase-3 concentrations are elevated.  相似文献   

2.
Crystallographic studies of Escherichia coli S-adenosylmethionine synthetase (ATP:L-methionine S-adenosyltransferase, MAT) have defined a flexible polypeptide loop that can gate access to the active site without contacting the substrates. The influence of the length and sequence of this active site loop on catalytic efficiency has been characterized in a mutant in which the E. coli MAT sequence (DRADPLEQ) has been replaced with the distinct sequence of the corresponding region of the otherwise highly homologous rat liver enzyme (HDLRNEEDV). Four additional mutants in which the entire DRADPLEQ sequence was replaced by five, six, seven, or eight glycines have been studied to unveil the effects of loop length and the influence of side chains. In all of the mutants, the maximal rate of S-adenosylmethionine formation (k(cat)) is diminished by more than 200-fold whereas the rate of hydrolysis of the tripolyphosphate intermediate is decreased by less than 3-fold. Thus, the function of the loop is localized to the first step in the overall reaction. The K(m) for methionine increases in all of the oligoglycine mutants, whereas the K(m) values for ATP are not substantially different. The k(cat) for the wild-type enzyme is decreased by increases in solution microviscosity with 55% of the maximal dependence. Thus, a diffusional event is coupled to the chemical step of AdoMet formation, which is known to be rate-limiting. The results indicate that a conformational change, possibly loop closure, is associated with AdoMet synthesis. The data integrate a previously discovered conformational change associated with PPP(i) binding to the E x AdoMet complex into the reaction sequence, reflecting a difference in protein conformation in the E x AdoMet x PPP(i) complex whether it is formed from the E x ATP x methionine complex or from binding of exogenous PPP(i). The temperature dependence of the k(cat) for S-adenosylmethionine formation shows that the removal of the side chains in the glycine mutants causes the activation enthalpy of the reaction to approximately double in each case, while the activation entropy changes from negative in the wild-type enzyme to positive in the mutants. The favorable activation entropy in the mutant-catalyzed reactions may reflect release of water during catalysis, while the negative activation entropy in the reaction catalyzed by the wild-type enzyme apparently reflects reorganization of the loop. The observations point to how nature can fine-tune the activity of an enzyme by modifying substrate and product access to the active site rather than by altering the enzyme x substrate contacts or the catalytic machinery itself.  相似文献   

3.
Tissue-type plasminogen activator (t-PA), the serine protease responsible for catalyzing the production of plasmin from plasminogen at the site of blood clots, is synthesized as a single-chain polypeptide precursor. Proteolytic cleavage at the C-terminal side of Arg275 generates a two-chain form of the enzyme whose subunits are held together by a single disulfide bond. We have measured the activities of both forms of the wild-type enzyme, as well as that of a mutant enzyme (Arg275----Gly), created by oligonucleotide-directed mutagenesis, that cannot be cleaved into a two-chain form. Both types of single-chain t-PAs are enzymatically active and exhibit identical Vmax and Km values when assayed with synthetic peptide substrates, indicating that the single amino acid change had no effect on the amidolytic activity of the enzyme. However, cleavage of wild-type t-PA into the two-chain form results in increased activity both on a peptide substrate and on the natural substrates Lys- and Glu-plasminogen in the absence or presence of stimulation by soluble fibrin. The enhanced activity is due to a 3-5-fold increase in the Vmax of the cleaved enzyme, rather than to any change in the Km values for the various substrates. During incubation with plasminogen, the single-chain form of wild-type t-PA is converted to the two-chain form by plasmin generated during the reaction. This conversion, from the less active form of the enzyme, results in a reaction that displays biphasic kinetics.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
5.
Nitric oxide (NO) signaling through the formation of cGMP is well established; however, there seems to be an increasing role for cGMP-independent NO signaling. Although key molecular details remain unanswered, S-nitrosation represents an example of cGMP-independent NO signaling. This modification has garnered recent attention as it has been shown to modulate the function of several important biochemical pathways. Although an analogy to O-phosphorylation can be drawn, little is known about protein nitrosothiol regulation in vivo. In solution, NO readily reacts with oxygen to yield a nitrosating agent, but this process alone provides no specificity for nitrosation. This lack of specificity is exemplified by the in vitro poly-S-nitrosation of caspase-3 (Casp-3, ref. 6) and the ryanodine receptor. Previous in vivo work with Casp-3 suggests that a protein-assisted process may be responsible for selective S-nitrosation of the catalytic cysteine (Cys163). We demonstrated that a single cysteine in thioredoxin (Trx) is capable of a targeted, reversible transnitrosation reaction with Cys163 of Casp-3. A greater understanding of how S-nitrosation is mediated has broad implications for cGMP-independent signaling. The example described here also suggests a new role for Trx in the regulation of apoptosis.  相似文献   

6.
A fragment of barnase comprising amino acids 1 to 36 (B(1-36)) that encompasses the region containing the two large helices (residues 6-18 and 26-34) of the native protein has been obtained by cleavage of the barnase mutant Val36----Met with cyanogen bromide. The circular dichroism (c.d.) spectrum of B(1-36) in the far ultraviolet indicates that the fragment is only weakly structured in water at neutral pH. The two-dimensional 1H nuclear magnetic resonance spectrum of B(1-36) shows, however, that a fraction of the population does have helical structure, spanning amino acid residues 8 to 18. B(1-36) becomes more helical in 35% trifluoroethanol. This is indicated by the c.d. spectrum and the increase from 6.6 to 7.0 in the pKa of His18, which is known to interact with the dipole of helix 6-18 in native barnase. The helical region of B(1-36) in 35% trifluoroethanol extends to residue 6. It is calculated from extrapolation of a trifluoroethanol titration of the ellipticity at 222 nm that B(1-36) exhibits in water approximately 6% of helical structure, calculated for a 36 residue alpha-helical peptide. This corresponds to approximately 20% of that expected for an 11-residue alpha-helical region. In trifluoroethanol, c.d. measurements indicate that approximately 30% of the 36-residue peptide is helical. It has been shown from extensive studies of the refolding of barnase that there is a folding intermediate that contains residues 8 to 18 in a helical conformation and that residue 6 is mainly unfolded. The experiments on the conformation of B(1-36) show that a small, but significant fraction, of its population in water adopts the conformation of the major alpha-helix during the barnase folding pathway, in the absence of tertiary interactions. Thus, in the folding of native barnase, secondary structure formation can precede the docking of the major alpha-helix onto the beta-sheet.  相似文献   

7.
Plasminogen activation by single-chain urokinase-type plasminogen activator or pro-urokinase (pro-UK) is accompanied by the generation of two-chain urokinase (UK) by plasmin which provides a positive feedback. In the present study, the time course of the activation of Glu-plasminogen and of Lys-plasminogen (10 microM) by pro-UK (1.0 nM) was studied. In the presence of native plasminogen (Glu-plasminogen), three distinct phases with different rates of plasmin generation were observed. The initial phase was slow and corresponded to the intrinsic activity of pro-UK as reflected by the activity of a plasmin-resistant mutant (Lys158----Ala). This was followed by a second phase which had the most rapid rate. The third phase had a plasminogen activation rate which was significantly slower than the second and paralleled the rate of activation by UK (1.0 nM). The second phase coincided with the time at which there was only about 50% conversion of pro-UK to UK, whereas the final phase coincided with essentially complete conversion. In the presence of fibrin fragment E-2 (20 microM), previously shown to strongly promote plasminogen activation by pro-UK, the identical phenomenon was observed, but at one-tenth the concentration of pro-UK. The most rapid rate of plasmin generation again coincided with transitional (25-60%) pro-UK to UK conversion. To further examine this phenomenon, the rate of pro-UK to UK conversion was controlled by using kallikrein in the presence of a plasmin inhibitor. In this experiment, the activation of Glu-plasminogen bound to solid-phase fibrin was measured. A similar three-phase sequence was observed, the highest rate of plasmin generation coinciding with about 45% conversion of pro-UK to UK. A mechanism for this transitional state phenomenon was postulated based on the established significantly higher affinity of pro-UK than of UK for Glu-plasminogen. This exceptional property for a proenzyme may enable a transient activity to be generated during the transition from pro-UK to UK corresponding to the more favorable KM of pro-UK and the kcat of UK. This hypothesis was supported by the results from experiments in which Lys-plasminogen was substituted for the Glu form. No transitional state activity was observed, consistent with the relatively high KM of pro-UK against Lys-plasminogen.  相似文献   

8.
Allingham JS  Sproul LR  Rayment I  Gilbert SP 《Cell》2007,128(6):1161-1172
Conventional kinesin and class V and VI myosins coordinate the mechanochemical cycles of their motor domains for processive movement of cargo along microtubules or actin filaments. It is widely accepted that this coordination is achieved by allosteric communication or mechanical strain between the motor domains, which controls the nucleotide state and interaction with microtubules or actin. However, questions remain about the interplay between the strain and the nucleotide state. We present an analysis of Saccharomyces cerevisiae Kar3/Vik1, a heterodimeric C-terminal Kinesin-14 containing catalytic Kar3 and the nonmotor protein Vik1. The X-ray crystal structure of Vik1 exhibits a similar fold to the kinesin and myosin catalytic head, but lacks an ATP binding site. Vik1 binds more tightly to microtubules than Kar3 and facilitates cooperative microtubule decoration by Kar3/Vik1 heterodimers, and yet allows motility. These results demand communication between Vik1 and Kar3 via a mechanism that coordinates their interactions with microtubules.  相似文献   

9.
Prolyl 4-hydroxylases (P4Hs) are 2-oxoglutarate dioxygenases that catalyze the hydroxylation of peptidyl prolines. They play an important role in collagen synthesis, oxygen homeostasis, and plant cell wall formation. We describe four structures of a P4H from the green alga Chlamydomonas reinhardtii, two of the apoenzyme at 1.93 and 2.90 A resolution, one complexed with the competitive inhibitor Zn2+, and one with Zn2+ and pyridine 2,4-dicarboxylate (which is an analogue of 2-oxoglutarate) at 1.85 A resolution. The structures reveal the double-stranded beta-helix core fold (jellyroll motif), typical for 2-oxoglutarate dioxygenases. The catalytic site is at the center of an extended shallow groove lined by two flexible loops. Mutagenesis studies together with the crystallographic data indicate that this groove participates in the binding of the proline-rich peptide-substrates. It is discussed that the algal P4H and the catalytic domain of collagen P4Hs have notable structural similarities, suggesting that these enzymes form a separate structural subgroup of P4Hs different from the hypoxia-inducible factor P4Hs. Key structural differences between these two subgroups are described. These studies provide first insight into the structure-function relationships of the collagen P4Hs, which unlike the hypoxia-inducible factor P4Hs use proline-rich peptides as their substrates.  相似文献   

10.
The catalytic domain of the peptidyl-prolyl cis/ trans isomerase Pin1 is a member of the FKBP superfold family. Within its active site are two highly conserved histidine residues, H59 and H157. Despite their sequence conservation in parvulin PPIase domains, the role of these histidine residues remains unclear. Our previous work (Behrsin et al. (2007) J. Mol. Biol. 365, 1143- 1162.) was consistent with a model where one or both histidines had critical roles in a hydrogen bonding network in the active site. Here, we test this model by looking at the effect of mutations to H59 and H157 on Pin1 function, activity, and protein stability. Using a yeast complementation assay, we show that both H59 and H157 can be mutated to non-hydrogen bonding residues and still support viability. Surprisingly, a nonfunctional H59L mutation can be rescued by a mutation of H157, to leucine. This double mutation (H59L/H157L) also had about 5-fold greater isomerase activity than the H59L mutation with a phosphorylated substrate. Structural analyses suggest that rescue of function and activity results from partial rescue of protein stability. Our findings indicate that H59 and H157 are not required for hydrogen bonding within the active site, and in contrast to the active site C113, they do not participate directly in catalysis. Instead, we suggest these histidines play a key role in domain structure or stability.  相似文献   

11.
Asp-60 is believed to be a catalytically essential residue of the tryptophan synthetase alpha chain of Escherichia coli (Nagata, S., Hyde, C.C., and Miles, E.W. (1989) J. Biol. Chem. 264, 6288-6296). Surprisingly, mutations altering Asp-60 were not observed in the many trpA missense mutants characterized in the 1960s. However, there was one genetic class of trpA missense mutants, represented by trpA34, for which protein structure analyses failed to detect an amino acid substitution. DNA sequence analyses have now shown that the trpA34 mutation was in codon 60 and that it resulted in replacement of Asp-60 by Asn. This finding provides additional support for the conclusion that the tryptophan synthetase alpha chain contains only a small number of absolutely essential residues.  相似文献   

12.
13.
The transition state of the group I intron self-splicing reaction is stabilized by three metal ions. The functional groups within the intron substrates (guanosine and an oligoribonucleotide mimic of the 5'-exon) that coordinate these metal ions have been systematically defined through a series of metal ion specificity switch experiments. In contrast, the catalytic metal ligands within the ribozyme active site are unknown. In an effort to identify them, stereospecific (R(P) or S(P)) single-site phosphorothioate substitutions were introduced at five phosphates predicted to be in the vicinity of the catalytic center (A207, C208, A304, U305, and A306) within the Tetrahymena intron. Of the 10 ribozymes that were studied, four phosphorothioate substitutions (A207 S(P), C208 S(P), A306 R(P), and A306 S(P)) exhibited a significant reduction in the cleavage rate. Only the effect of the C208 S(P) phosphorothioate substitution could be significantly rescued by the addition of a thiophilic metal ion, either Mn(2+) or Zn(2+), when tested with an all-oxy substrate. The effect was not rescued with Cd(2+). To determine if one of the catalytic metal ions is coordinated to the C208 pro-S(P) oxygen, the phosphorothioate-substituted ribozymes were also assayed using oligonucleotide substrates with a 3'-phosphorothiolate or an S(P) phosphorothioate substitution at the scissile phosphate. This resulted in a second metal specificity switch, in that Mn(2+) or Zn(2+) no longer rescued the C208 S(P) ribozyme, but Cd(2+) provided efficient rescue in the context of either sulfur-containing substrate. The 3'-oxygen and the pro-S(P) oxygen of the scissile phosphate are both known to coordinate the same metal ion, M(A), which stabilizes the negative charge on the leaving group 3'-oxygen in the transition state. Taken together, these data suggest that metal M(A) is coordinated to the C208 pro-S(P) phosphate oxygen, which constitutes the first functional link between a specific catalytic metal ion and a particular functional group within the group I ribozyme active site.  相似文献   

14.
Factor (F) VIII functions as a cofactor in FXase, markedly accelerating the rate of FIXa-catalyzed activation of FX. Earlier work identified a FX-binding site having μM affinity within the COOH-terminal region of the FVIIIa A1 subunit. In the present study, surface plasmon resonance (SPR), ELISA-based binding assays, and chemical cross-linking were employed to assess an interaction between FX and the FVIII light chain (A3C1C2 domains). SPR and ELISA-based assays showed that FVIII LC bound to immobilized FX (K(d) = 165 and 370 nM, respectively). Furthermore, active site-modified activated protein C (DEGR-APC) effectively competed with FX in binding FVIII LC (apparent K(i) = 82.7 nM). Western blotting revealed that the APC-catalyzed cleavage rate at Arg(336) was inhibited by FX in a concentration-dependent manner. A synthetic peptide comprising FVIII residues 2007-2016 representing a portion of an APC-binding site blocked the interaction of FX and FVIII LC (apparent K(i) = 152 μM) and directly bound to FX (K(d) = 7.7 μM) as judged by SPR and chemical cross-linking. Ala-scanning mutagenesis of this sequence revealed that the A3C1C2 subunit derived from FVIII variants Thr2012Ala and Phe2014Ala showed 1.5- and 1.8-fold increases in K(d) for FX, whereas this value using the A3C1C2 subunit from a Thr2012Ala/Leu2013Ala/Phe2014Ala triple mutant was increased >4-fold. FXase formed using this LC triple mutant demonstrated an ~4-fold increase in the K(m) for FX. These results identify a relatively high affinity and functional FX site within the FVIIIa A3C1C2 subunit and show a contribution of residues Thr2012 and Phe2014 to this interaction.  相似文献   

15.
Histidine 228 at the active site of Escherichia coli serine hydroxymethyltransferase was replaced with an asparagine. The mutant enzyme was expressed in a strain of E. coli that lacks wild type enzyme. Absorption spectra, circular dichroism spectra, and differential scanning calorimetry thermograms suggest that the amino acid change at the active site causes no detectable change in the tertiary structure of the enzyme. Kinetic studies demonstrated that kcat for the mutant enzyme is about 25% of the value for the wild type enzyme with either L-serine or allothreonine as substrate. Km or Kd values for amino acid substrates and reduced folate compounds were 2-10-fold larger with the mutant enzyme. The rate of interconversion of several enzyme-glycine complexes showed that the conversion of the external aldimine to the quinoid complex is not the rate-determining step for either the mutant or wild type enzyme in the presence of tetrahydrofolate. The binding of L-serine to the wild type enzyme gives a more thermally stable enzyme and increases its affinity for tetrahydrofolate. These effects are not found when L-serine binds to the mutant enzyme. The studies demonstrate that histidine 228 is not a catalytically essential residue and suggest that it is involved in interacting with either the amino acid substrate or the enzyme-bound pyridoxal phosphate.  相似文献   

16.
ATP-dependent Lon protease degrades specific short-lived regulatory proteins as well as defective and abnormal proteins in the cell. The crystal structure of the proteolytic domain (P domain) of the Escherichia coli Lon has been solved by single-wavelength anomalous dispersion and refined at 1.75-A resolution. The P domain was obtained by chymotrypsin digestion of the full-length, proteolytically inactive Lon mutant (S679A) or by expression of a recombinant construct encoding only this domain. The P domain has a unique fold and assembles into hexameric rings that likely mimic the oligomerization state of the holoenzyme. The hexamer is dome-shaped, with the six N termini oriented toward the narrower ring surface, which is thus identified as the interface with the ATPase domain in full-length Lon. The catalytic sites lie in a shallow concavity on the wider distal surface of the hexameric ring and are connected to the proximal surface by a narrow axial channel with a diameter of approximately 18 A. Within the active site, the proximity of Lys(722) to the side chain of the mutated Ala(679) and the absence of other potential catalytic side chains establish that Lon employs a Ser(679)-Lys(722) dyad for catalysis. Alignment of the P domain catalytic pocket with those of several Ser-Lys dyad peptide hydrolases provides a model of substrate binding, suggesting that polypeptides are oriented in the Lon active site to allow nucleophilic attack by the serine hydroxyl on the si-face of the peptide bond.  相似文献   

17.
Badarau A  Page MI 《Biochemistry》2006,45(35):10654-10666
The kinetics and mechanism of hydrolysis of the native zinc and metal substituted Bacillus cereus (BcII) metallo-beta-lactamase have been investigated. The pH and metal ion dependence of k(cat) and k(cat)/K(m), determined under steady-state conditions, for the cobalt substituted BcII catalyzed hydrolysis of cefoxitin, cephaloridine, and cephalexin indicate that an enzyme residue of apparent pK(a) 6.3 +/- 0.1 is required in its deprotonated form for metal ion binding and catalysis. The k(cat)/K(m) for cefoxitin and cephalexin with cadmium substituted BcII is dependent on two ionizing groups on the enzyme: one of pK(a1) = 8.7 +/- 0.1 required in its deprotonated form and the other of pK(a2) = 9.3 +/- 0.1 required in its protonated form for activity. The pH dependence of the competitive inhibition constant, K(i), for CdBcII with l-captopril indicates that pK(a1) = 8.7 +/- 0.1 corresponds to the cadmium-bound water. For the manganese substituted BcII, the pH dependence of k(cat)/K(m) for benzylpenicillin, cephalexin, and cefoxitin similarly indicated the importance of two catalytic groups: one of pK(a1) = 8.5 +/- 0.1 which needs to be deprotonated and the other of pK(a2) = 9.4 +/- 0.1 which needs to be protonated for catalysis; the pK(a1) was assigned to the manganese-bound water. The rate was metal ion concentration dependent at the highest manganese concentrations used (10(-)(3) M). The metal substituted species have similar or higher catalytic activities compared with the zinc enzyme, albeit at pHs above 7. Interestingly, with cefoxitin, a very poor substrate for ZnBcII, both k(cat) and k(cat)/K(m) increase with increasing pK(a) of the metal-bound water, in the order Zn < Co < Mn < Cd. A higher pK(a) for the metal-bound water for cadmium and manganese BCII leads to more reactive enzymes than the native zinc BcII, suggesting that the role of the metal ion is predominantly to provide the nucleophilic hydroxide, rather than to act as a Lewis acid to polarize the carbonyl group and stabilize the oxyanion tetrahedral intermediate.  相似文献   

18.
19.
A revertant was isolated from a temperature-sensitive poliovirus 2C mutant, 2C-31, which is defective in viral RNA synthesis. This revertant, called 2C-31R1, grew well at 39 degrees C and was not defective in RNA synthesis. However, in contrast to its parental mutant, 2C-31R1 was cold sensitive and could hardly grow at all at 32 degrees C. Analysis of a single-cycle growth revealed that 2C-31R1 was defective in virion uncoating at 32 degrees C, and a substantial amount (more than 30%) of input viruses could be recovered as infectious particles from an infected cell lysate up to 6 h postinfection. The uncoating defect and the inability to grow at cold temperatures could be overcome by a brief incubation at the permissive temperature (39 degrees C) before the infection was continued at 32 degrees C. cDNA cloning and mix-and-match recombination experiments indicated that the defect in uncoating was the result of two secondary point mutations, seven nucleotides apart, in the 2C-coding sequence downstream of the inserted linker which is the original mutation in the parental 2C-31 genome. Another revertant, 2C-31R3, isolated from the same 2C-31 stock, was not defective in uncoating and appeared to be a secondary revertant that contained an intragenic suppressor for the uncoating defect. The uncoating defect of 2C-31R1 could be complemented by type 2 poliovirus. These results suggested that protein 2C, in addition to its role in viral RNA synthesis, has a function in determining virion structure.  相似文献   

20.
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