首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Oxidation of the soluble, truncated form of cytochrome f by wild-type and mutant species of plastocyanin has been analyzed by laser flash absorption spectroscopy in the cyanobacterium Nostoc (formerly, Anabaena) sp. PCC 7119. At low ionic strengths, the apparent electron transfer rate constant of cytochrome f oxidation by wild-type plastocyanin is 1.34 x 10(4) s(-)(1), a value much larger than those determined for the same proteins from other organisms. Upon site-directed mutagenesis of specific residues at the plastocyanin interaction area, the rate constant decreases in all cases yet to varying extents. The only exception is the D54K variant, which exhibits a higher reactivity toward cytochrome f. In most cases, the reaction rate constant decreases monotonically with an increase in ionic strength. The observed changes in the reaction mechanism and rate constants are in agreement with the location of the mutated residues at the interface area, as well as with the peculiar orientation of the two partners within the Nostoc plastocyanin-cytochrome f transient complex, whose NMR structure has been determined recently. Furthermore, the experimental data herein reported match well the kinetic behavior exhibited by the same set of plastocyanin mutants when acting as donors of electrons to photosystem I [Molina-Heredia, F. P., et al. (2001) J. Biol. Chem. 276, 601-605], thus indicating that the copper protein uses the same surface areas-one hydrophobic and the other electrostatic-to interact with both cytochrome f and photosystem I.  相似文献   

2.
The X-ray crystal structures of oxidized and reduced plastocyanin obtained from the fern Dryopteris crassirhizoma have been determined at 1.7 and 1.8 A resolution, respectively. The fern plastocyanin is unique in the longer main chain composed of 102 amino acid residues and in the unusual pH dependence due to the pi-pi stacking interaction around the copper site [Kohzuma, T., et al. (1999) J. Biol. Chem. 274, 11817-11823]. Here we report the structural comparison between the fern plastocyanin and other plastocyanins from cyanobacteria, green algae, and other higher plants, together with the structural changes of fern plastocyanin upon reduction. Glu59 hydrogen bonds to the OH of Tyr83, which is thought to be a possible conduit for electrons, in the oxidized state. However, it moves away from Tyr83 upon reduction like poplar plastocyanin.  相似文献   

3.
Three surface residues of plastocyanin from Prochlorothrix hollandica have been modified by site-directed mutagenesis. Changes have been made in methionine 33, located in the hydrophobic patch of the copper protein, and in arginine 86 and proline 53, both located in the eastern hydrophilic area. The reactivity toward photosystem I of single mutants M33N, P53A, P53E, R86Q, R86E, and the double mutant M33N/P14L has been studied by laser flash absorption spectroscopy. All the mutations yield increased reactivity of plastocyanin toward photosystem I as compared with wild type plastocyanin, thus indicating that in Prochlorothrix electron donation to photosystem I is not optimized. The most drastic increases in the intracomplex electron transfer rate are obtained with mutants in methionine 33, whereas replacing arginine 86 only modestly affects the plastocyanin-photosystem I equilibrium constant for complex formation. Mutations at position 53 also promote major changes in the association of plastocyanin with photosystem I, yielding a change from a mechanism involving complex formation to a simpler collisional interaction. Molecular dynamics calculations indicate that mutations at position 33 promote changes in the H-bond network around the copper center. The comparative kinetic analysis of the reactivity of Prochlorothrix plastocyanin mutants toward photosystem I from other cyanobacteria reveals that mutations M33N, P53A, and P53E result in enhanced general reactivity.  相似文献   

4.
The two-pore (2P) domain K(+) channels TREK-1 and TRAAK are opened by membrane stretch as well as arachidonic acid (AA) (Patel, A. J., Honoré, E., Maingret, F., Lesage, F., Fink, M., Duprat, F., and Lazdunski, M. (1998) EMBO J. 17, 4283-4290; Maingret, F., Patel, A. J., Lesage, F., Lazdunski, M., and Honoré, E. (1999) J. Biol. Chem. 274, 26691-26696; Maingret, F., Fosset, M., Lesage, F., Lazdunski, M. , and Honoré, E. (1999) J. Biol. Chem. 274, 1381-1387. We demonstrate that lysophospholipids (LPs) and platelet-activating factor also produce large specific and reversible activations of TREK-1 and TRAAK. LPs activation is a function of the size of the polar head and length of the acyl chain but is independent of the charge of the molecule. Bath application of lysophosphatidylcholine (LPC) immediately opens TREK-1 and TRAAK in the cell-attached patch configuration. In excised patches, LPC activation is lost, whereas AA still produces maximal opening. The carboxyl-terminal region of TREK-1, but not the amino terminus and the extracellular loop M1P1, is critically required for LPC activation. LPC activation is indirect and may possibly involve a cytosolic factor, whereas AA directly interacts with either the channel proteins or the bilayer and mimics stretch. Opening of TREK-1 and TRAAK by fatty acids and LPs may be an important switch in the regulation of synaptic function and may also play a protective role during ischemia and inflammation.  相似文献   

5.
Chan BS  Bao Y  Schuster VL 《Biochemistry》2002,41(29):9215-9221
The prostaglandin transporter "PGT" interacts electrostatically with its anionic substrate, based on inhibition by the disulfonic stilbenes [Chan, B. S. (1998) J. Biol. Chem. 273, 6689-6697], inhibition by the thiol-reactive anion sodium (2-sulfonatoethyl)methanethiosulfonate (MTSES) [Chan, B. S. (1999) J. Biol. Chem. 274, 25564-25570], and the requirement for a negatively charged 1-position carboxyl on the substrate [Itoh, S. (1996) Mol. Pharm. 50, 736-742]. Here we found that modification of positively charged residues on wild-type PGT by arginine- and lysine-specific reagents significantly inhibited transport. We previously found that the binding site of PGT is formed, at least in part, by its membrane-spanning segments [Chan, B. S. (1999) J. Biol. Chem. 274, 25564-25570]. Three charged residues within predicted transmembrane spans (E78, R560, and K613) are conserved in PGT and in related transporters. Substitution of the anionic residue E78 (E78D and E78C) produced an essentially functional transporter, whereas substitution of the cationic residues with neutral residues (R560N and K613Q) resulted in poorly functional transporters. Immunoblotting revealed similar expression levels of wild-type and mutant transporters, and immunostaining indicated correct targeting. Conservative charge substitutions (R560K, K613R, and K613H) resulted in generally functional transporters. In contrast, R560N was nonfunctional, whereas the substrate affinity of K613G decreased greater than 50-fold. Conservative substitutions retaining the charge at position 613 (K613R and K613H) restored the substrate affinity, suggesting a direct role of K613 in substrate binding. Double-neutral mutants E78G/R560C and E78G/K613C were inactive, indicating that these residues are not simply charge-paired. Our results suggest that an arginine at position 560 is critical for maximal substrate translocation, and that a positively charged side chain at position 613 contributes to electrostatic binding of the anionic substrate.  相似文献   

6.
Long chain curarimimetic toxins from snake venom bind with high affinities to both muscular type nicotinic acetylcholine receptors (AChRs) (K(d) in the pm range) and neuronal alpha 7-AChRs (K(d) in the nm range). To understand the molecular basis of this dual function, we submitted alpha-cobratoxin (alpha-Cbtx), a typical long chain curarimimetic toxin, to an extensive mutational analysis. By exploring 36 toxin mutants, we found that Trp-25, Asp-27, Phe-29, Arg-33, Arg-36, and Phe-65 are involved in binding to both neuronal and Torpedo (Antil, S., Servent, D., and Ménez, A. (1999) J. Biol. Chem. 274, 34851-34858) AChRs and that some of them (Trp-25, Asp-27, and Arg-33) have similar binding energy contributions for the two receptors. In contrast, Ala-28, Lys-35, and Cys-26-Cys-30 selectively bind to the alpha 7-AChR, whereas Lys-23 and Lys-49 bind solely to the Torpedo AChR. Therefore, alpha-Cbtx binds to two AChR subtypes using both common and specific residues. Double mutant cycle analyses suggested that Arg-33 in alpha-Cbtx is close to Tyr-187 and Pro-193 in the alpha 7 receptor. Since Arg-33 of another curarimimetic toxin is close to the homologous alpha Tyr-190 of the muscular receptor (Ackermann, E. J., Ang, E. T. H., Kanter, J. R., Tsigelny, I., and Taylor, P. (1998) J. Biol. Chem. 273, 10958-10964), toxin binding probably occurs in homologous regions of neuronal and muscular AChRs. However, no coupling was seen between alpha-Cbtx Arg-33 and alpha 7 receptor Trp-54, Leu-118, and Asp-163, in contrast to what was observed in a homologous situation involving another toxin and a muscular receptor (Osaka, H., Malany, S., Molles, B. E., Sine, S. M., and Taylor, P. (2000) J. Biol. Chem. 275, 5478-5484). Therefore, although occurring in homologous regions, the detailed modes of toxin binding to alpha 7 and muscular receptors are likely to be different. These data offer a molecular basis for the design of toxins with predetermined specificities for various members of the AChR family.  相似文献   

7.
The present studies demonstrate that no single stretch of sequence in the third intracellular (3i) loop of the alpha(2A) adrenergic receptor (alpha(2A)-AR) can fully account for its previously described interactions with spinophilin (Richman, J. G., Brady, A. E., Wang, Q., Hensel, J. L., Colbran, R. J., and Limbird, L. E. (2001) J. Biol. Chem. 276, 15003-15008), 14-3-3zeta (Prezeau, L., Richman, J. G., Edwards, S. W., and Limbird, L. E. (1999) J. Biol. Chem. 274, 13462-13469), and arrestin 3 (Wu, G., Krupnick, J. G., Benovic, J. L., and Lanier, S. M. (1997) J. Biol. Chem. 272, 17836-17842), suggesting that a three-dimensional surface, rather than a linear sequence, provides the basis for these interactions as proposed for 3i loop tethering of the alpha(2A)-AR to the basolateral surface of Madin-Darby canine kidney cells (Edwards, S. W., and Limbird, L. E. (1999) J. Biol. Chem. 274, 16331-16336). Sequences at the extreme N-terminal and C-terminal ends of the 3i loop are critical for interaction with spinophilin but not for interaction with 14-3-3zeta or arrestin 3, for which the C-terminal half of the loop is more important. Competition binding for (35)S-labeled alpha(2A)-AR 3i loop binding to glutathione S-transferase (GST)-spinophilin amino acids 151-444 revealed a relative affinity of spinophilin congruent with arrestin > 14-3-3zeta for the unphosphorylated alpha(2A)-AR 3i loop. Agonist occupancy of the alpha(2A)-AR increases receptor association with spinophilin, and arrestin 3 appears to compete for this enrichment. However, when the G protein-coupled receptor kinase 2 substrate sequence was deleted from the 3i loop, arrestin 3 could not compete for the agonist-enriched binding of spinophilin to the mutant alpha(2A)-AR. These data are consistent with a model where sequential or competitive interactions among spinophilin, arrestin, and/or 14-3-3zeta play a role in alpha(2A)-AR functions.  相似文献   

8.
9.
To determine the interaction site for cytochrome c (Cc) on cytochrome c oxidase (CcO), a number of conserved carboxyl residues in subunit II of Rhodobacter sphaeroides CcO were mutated to neutral forms. A highly conserved tryptophan, Trp(143), was also mutated to phenylalanine and alanine. Spectroscopic and metal analyses of the surface carboxyl mutants revealed no overall structural changes. The double mutants D188Q/E189N and D151Q/E152N exhibit similar steady-state kinetic behavior as wild-type oxidase with horse Cc and R. sphaeroides Cc(2), showing that these residues are not involved in Cc binding. The single mutants E148Q, E157Q, D195N, and D214N have decreased activities and increased K(m) values, indicating they contribute to the Cc:CcO interface. However, their reactions with horse and R. sphaeroides Cc are different, as expected from the different distribution of surface lysines on these cytochromes c. Mutations at Trp(143) severely inhibit activity without changing the K(m) for Cc or disturbing the adjacent Cu(A) center. From these data, we identify a Cc binding area on CcO with Trp(143) and Asp(214) close to the site of electron transfer and Glu(148), Glu(157), and Asp(195) providing electrostatic guidance. The results are completely consistent with time-resolved kinetic measurements (Wang, K., Zhen, Y., Sadoski, R., Grinnell, S., Geren, L., Ferguson-Miller, S., Durham, B., and Millett, F. (1999) J. Biol. Chem. 274, 38042-38050) and computational docking analysis (Roberts, V. A., and Pique, M. E. (1999) J. Biol. Chem. 274, 38051-38060).  相似文献   

10.
The enamel protein amelogenin binds to GlcNAc (Ravindranath, R. M. H., Moradian-Oldak, R., and Fincham, A.G. (1999) J. Biol. Chem. 274, 2464-2471) and to the GlcNAc-mimicking peptide (GMp) (Ravindranath, R. M. H., Tam, W., Nguyen, P., and Fincham, A. G. (2000) J. Biol. Chem. 275, 39654-39661). The GMp motif in the N-terminal region of the cytokeratin 14 of ameloblasts binds to trityrosyl motif peptide (ATMP) of amelogenin (Ravindranath, R. M. H., Tam, W., Bringas, P., Santos, V., and Fincham, A. G. (2001) J. Biol. Chem. 276, 36586 - 36597). K14 (Type I) pairs with K5 (Type II) in basal epithelial cells; GlcNAc-acylated K5 is identified in ameloblasts. Dosimetric analysis showed the binding affinity of amelogenin to K5 and to GlcNAc-acylated-positive control, ovalbumin. The specific binding of [3H]ATMP with K5 or ovalbumin was confirmed by Scatchard analysis. [3H]ATMP failed to bind to K5 after removal of GlcNAc. Blocking K5 with ATMP abrogates the K5-amelogenin interaction. K5 failed to bind to ATMP when the third proline was substituted with threonine, as in some cases of human X-linked amelogenesis imperfecta or when tyrosyl residues were substituted with phenylalanine. Confocal laser scan microscopic observations on ameloblasts during postnatal (PN) growth of the teeth showed that the K5-amelogenin complex migrated from the cytoplasm to the periphery (on PN day 1) and accumulated at the apical region on day 3. Secretion of amelogenin commences from day 1. K5, similar to K14, may play a role of chaperone during secretion of amelogenin. Upon secretion of amelogenin, K5 pairs with K14. Pairing of K5 and K14 commences on day 3 and ends on day 9. The pairing of K5 and K14 marks the end of secretion of amelogenin.  相似文献   

11.
12.
A C Anusiem  M Kelleher 《Biopolymers》1984,23(7):1147-1167
Interest in the thermodynamics of the iron-binding site in hemoproteins has increased in recent years due to refinements in x-ray crystallographic studies of hemoproteins [see Deathage, J. F., Lee, R. S., Anderson, C. M. & Moffat, K. (1976) J. Mol. Biol. 104 , 687–706; Heidner, E. J., Ladner, R. C. & Perutz, M. F. (1976) J. Mol. Biol. 104 , 707–722; Deathage, J. F., Lee, R. S. & Moffat, K. (1976) J. Mol. Biol. 104 , 723–728; Ladner, R. C., Heidner, E. J. & Perutz, M. F. (1976) J. Mol. Biol. 114 , 385–414; Fermi, G. & Perutz, M. F. (1977) J. Mol. Biol. 114 , 421–431; Takano, T. (1977) J. Mol. Biol. 110 , 537–568 and 569–589], the synthesis and x-ray analysis of model heme compounds [see Scheidt, W. R. (1977) Acc. Chem. Res. 10 , 339–345; Kastner, M. E., Scheidt, W. R., Mashino, T. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 666–667; Mashiko, T., Kastner, M. E., Spartalian, K., Scheidt, W. R. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 6354–6362; Hill, H. A. O., Skite, P. P., Buchler, J. W., Luchr, H., Tonn, M., Gregson, A. K. & Pellizer, G. (1979) Chem. Commun. 4 , 151–152; and Scheidt, W. R., Cohen, I. A. & Kastner, M. E. (1979) Biochemistry 18 , 3546–3556], and the numerous data on heme–protein interactions that account for the differences observed in ligand binding between the various species of animals. Numerous probes have been used and provide information about the structure and thermodynamics of the binding site, but no single probe can provide the complete picture [see Iizuka, T. & Yonetani, T. (1970) Adv. Biophys. 1 , 157–182; Smith, D. W. & Williams, R. J. P. (1970) Struct. Bond. 7 , 1–45; and Spiro, T. G. (1975) Biochim. Biophys. Acta 416 , 169–189].  相似文献   

13.
Asn792 present in M5 of gastric H,K-ATPase is highly conserved within the P-type ATPase family. A direct role in K+ binding was postulated for Na,K-ATPase but was not found in a recent model for gastric H,K-ATPase (Koenderink, J. B., Swarts, H. G. P., Willems, P. H. G. M., Krieger, E., and De Pont, J. J. H. H. M. (2004) J. Biol. Chem. 279, 16417-16424). Therefore, its role in K+ binding and E1/E2 conformational equilibrium in gastric H,K-ATPase was studied by site-directed mutagenesis and expression in Sf9 cells. N792Q and N792A, but not N792D and N792E, had a markedly reduced K+ affinity in both the ATPase and dephosphorylation reactions. In addition, N792A shifted the conformational equilibrium to the E1 form. In double mutants, the effect of N792A on K+ sensitivity was overruled by either E820Q (K(+)-independent activity) or E343D (no dephosphorylation activity). Models were made for the mutants based on the E2 structure of Ca(2+)-ATPase. In the wild-type model the acid amide group of Asn792 has hydrogen bridges to Lys791, Ala339, and Val341. Comparison of the effects of the various mutants suggests that the hydrogen bridge between the carbonyl oxygen of Asn792 and the amino group of Lys791 is essential for the K+ sensitivity and the E2 preference of wild-type enzyme. Moreover, there was a high positive correlation (r = 0.98) between the in silico calculated energy difference of the E2 form (mutants versus wild type) and the experimentally measured IC50 values for vanadate, which reflects the direction of the E2<-->E1 conformational equilibrium. These data strongly support the validity of the model in which Asn792 participates in the hydrogen bond network around the K(+)-binding pocket.  相似文献   

14.
15.
The reaction between cytochrome c (Cc) and Rhodobacter sphaeroides cytochrome c oxidase (CcO) was studied using a cytochrome c derivative labeled with ruthenium trisbipyridine at lysine 55 (Ru-55-Cc). Flash photolysis of a 1:1 complex between Ru-55-Cc and CcO at low ionic strength results in electron transfer from photoreduced heme c to Cu(A) with an intracomplex rate constant of k(a) = 4 x 10(4) s(-1), followed by electron transfer from Cu(A) to heme a with a rate constant of k(b) = 9 x 10(4) s(-1). The effects of CcO surface mutations on the kinetics follow the order D214N > E157Q > E148Q > D195N > D151N/E152Q approximately D188N/E189Q approximately wild type, indicating that the acidic residues Asp(214), Glu(157), Glu(148), and Asp(195) on subunit II interact electrostatically with the lysines surrounding the heme crevice of Cc. Mutating the highly conserved tryptophan residue, Trp(143), to Phe or Ala decreased the intracomplex electron transfer rate constant k(a) by 450- and 1200-fold, respectively, without affecting the dissociation constant K(D). It therefore appears that the indole ring of Trp(143) mediates electron transfer from the heme group of Cc to Cu(A). These results are consistent with steady-state kinetic results (Zhen, Y., Hoganson, C. W., Babcock, G. T., and Ferguson-Miller, S. (1999) J. Biol. Chem. 274, 38032-38041) and a computational docking analysis (Roberts, V. A., and Pique, M. E. (1999) J. Biol. Chem. 274, 38051-38060).  相似文献   

16.
The DNA duplex corresponding to the entire length (126 nucleotides) of the precursor for an Escherichia coli tyrosine tRNA has been synthesized. Duplex [I] (Sekiya, T., Besmer, P., Takeya, T., and Khorana, H. G.(1976) J. Biol. Chem. 251, 634-641), corresponding to the nucleotide sequence 1-26, containing single-stranded ends and carrying one appropriately labeled 5'-phosphate group, was joined to duplex [II] (Loewen, P. C., Miller, R. C., Panet, A., Sekiya, T., and Khorana, H. G. (1976) J. Biol. Chem. 251, 642-650) (nucleotide sequence 23-66 or 23-60) was phosphorylated with [gamma-33P]ATP at the 5'-OH ends. Duplex [III] (Panet, A., Kleppe, R., Kleppe, K., and Khorana, H. G. (1976) J. Biol. Chem. 251, 651-657) (nucleotide sequence 57-94 (Fig. 2)) was also phosphorylated at 5'-ends with [gamma-33P]ATP and was joined to duplex [IV] (Caruthers, M. H., Kleppe, R., Kleppe, K., and Khorana, H. G. (1976) J. Biol. Chem. 251, 658-666) (nucleotide sequence 90-126) which carried a 33P-labeled phosphate group on nucleotide 90. The joined product, duplex [III + IV] (nucleotide sequence 57-126) was characterized. The latter duplex was joined to the duplex [I + II] to give the total duplex. The latter contains singlestranded ends (nucleotides 1 to 6 and 121 to 126) which can either be "filled in" to produce the completely base-paired duplex or may be used to add the promoter and terminator regions at the appropriate ends.  相似文献   

17.
Agonist-elicited receptor sequestration is strikingly different for the alpha(2A)- versus alpha(2B)-adrenergic receptor (alpha(2)-AR) subtypes; the alpha(2B)-AR undergoes rapid and extensive disappearance from the HEK 293 cell surface, whereas the alpha(2A)-AR does not (Daunt, D. A., Hurt, C., Hein, L., Kallio, J., Feng, F., and Kobilka, B. K. (1997) Mol. Pharmacol. 51, 711-720; Eason, M. G., and Liggett, S. B. (1992) J. Biol. Chem. 267, 25473-25479). Since recent reports suggest that endocytosis is required for some G protein-coupled receptors to stimulate the mitogen-activated protein (MAP) kinase cascade (Daaka, Y., Luttrell, L. M., Ahn, S., Della Rocca, G. J., Ferguson, S. S., Caron, M. G., and Lefkowitz, R. J. (1998) J. Biol. Chem. 273, 685-688; Luttrell, L. M., Daaka, Y., Della Rocca, G. J., and Lefkowitz, R. J. (1997) J. Biol. Chem. 272, 31648-31656; Ignatova, E. G., Belcheva, M. M., Bohn, L. M., Neuman, M. C., and Coscia, C. J. (1999) J. Neurosci. 19, 56-63), we evaluated the differential ability of these two subtypes to activate MAP kinase. We observed no correlation between subtype-dependent agonist-elicited receptor redistribution and receptor activation of the MAP kinase cascade. Furthermore, incubation of cells with K(+)-depleted medium eliminated alpha(2B)-AR internalization but did not eliminate MAP kinase activation, suggesting that receptor internalization is not a general prerequisite for activation of the MAP kinase cascade via G(i)-coupled receptors. We also noted that neither dominant negative dynamin (K44A) nor concanavalin A treatment dramatically altered MAP kinase activation or receptor redistribution, indicating that these experimental tools do not universally block G protein-coupled receptor internalization.  相似文献   

18.
We have previously discovered the opium alkaloid noscapine as a microtubule interacting agent that binds to tubulin, alters the dynamics of microtubule assembly, and arrests mammalian cells at mitosis (Ye, K., Ke, Y., Keshava, N., Shanks, J., Kapp, J. A., Tekmal, R. R., Petros, J., and Joshi, H. C. (1998) Proc. Natl. Acad. Sci. U. S. A. 95, 1601-1606; Ye, K., Zhou, J., Landen, J. W., Bradbury, E. M., and Joshi, H. C. (2001) J. Biol. Chem. 276, 46697-46700; Zhou, J., Panda, D., Landen, J. W., Wilson, L., and Joshi, H. C. (2002) J. Biol. Chem. 277, 17200-17208). Here we show that noscapine does not compete with paclitaxel for tubulin binding and can efficiently inhibit the proliferation of both paclitaxel-sensitive and paclitaxel-resistant human ovarian carcinoma cells (i.e. the parental cell line 1A9 and two derivative cell lines, 1A9PTX10 and 1A9PTX22, which harbor beta-tubulin mutations that impair paclitaxel-tubulin interaction (Giannakakou, P., Sackett, D. L., Kang, Y. K., Zhan, Z., Buters, J. T., Fojo, T., and Poruchynsky, M. S. (1997) J. Biol. Chem. 272, 17118-17125). Strikingly, these cells undergo apoptotic death upon noscapine treatment, accompanied by activation of the c-Jun NH(2)-terminal kinases (JNK). Furthermore, inhibition of JNK activity by treatment with antisense oligonucleotide or transfection with dominant-negative JNK blocks noscapine-induced apoptosis. These findings thus indicate a great potential for noscapine in the treatment of paclitaxel-resistant human cancers. In addition, our results suggest that the JNK pathway plays an essential role in microtubule inhibitor-induced apoptosis.  相似文献   

19.
Previous crystallographic studies of the antibacterial trimethoprim in complexes with bacterial and avian dihydrofolate reductases have shown substantial differences in the mode of binding, providing plausible explanations for the origin of the remarkable species selectivity of this inhibitor (Matthews, D. A., Bolin, J. T., Burridge, J. M., Filman, D. J., Volz, K. W., Kaufman, B. T., Beddell, C. R., Champness, J. N., Stammers, D. K., and Kraut, J. (1985) J. Biol. Chem. 260, 381-391; Matthews, D. A., Bolin, J. T., Burridge, J. M., Filman, D. J., Volz, K. W., and Kraut, J. (1985) J. Biol. Chem. 260, 392-399). A major species difference between the active sites is that the only carboxylate present is always Glu in vertebrates and Asp in bacteria. Crystallographic studies of the wild-type and E30D mutant of the enzyme from mouse now reveal that in both cases trimethoprim is bound in an identical fashion to that observed with the bacterial enzyme, and there is no obvious single explanation for the origin of the 10(5)-fold selectivity of trimethoprim binding. In an earlier study of a mouse wild-type enzyme using more limited data it was proposed that trimethoprim bound in the avian mode (Stammers, D. K., Champness, J. N., Beddell, C. R., Dann, J. G., Eliopoulos, E. E., Geddes, A. J., Ogg, D., and North, A. C. T. (1987) FEBS Lett. 218, 178-184), but a re-examination indicates that the occupancy of the active site by trimethoprim is less than had been thought, and we are currently unable to make an unambiguous interpretation of the electron density maps and cannot confirm the avian mode of binding in those crystals.  相似文献   

20.
We report the isolation and analysis of genomic clones containing the entire gene encoding the muscle-specific subunit of human phosphoglycerate mutase. The gene spans 2.83 kilobase pairs and has a three-exon/two-intron structure that is similar to the organization of the human 2,3-bisphosphoglycerate mutase gene (Joulin, V., Garel, M.-C., LeBoulch, P., Valentin, C., Rosa, R., Rosa, J., and Cohen-Solal, M. (1988) J. Biol. Chem. 263, 15785-15790), in that the second introns of both genes are localized precisely at the same position. A canonical TATA box and an inverted CCAAT box are present immediately upstream of this gene. Comparison with other muscle-specific enzyme genes reveals a conserved 9-base pair element (GGGGCTGGG) in the 5'-flanking region that may be associated with the expression of genes encoding muscle-specific enzymes.  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