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 265 毫秒
1.
Residual dipolar couplings (RDCs) are NMR parameters that provide both structural and dynamic information concerning inter-nuclear vectors, such as N–HN and Cα–Hα bonds within the protein backbone. Two approaches for extracting this information from RDCs are the model free analysis (MFA) (Meiler et al. in J Am Chem Soc 123:6098–6107, 2001; Peti et al. in J Am Chem Soc 124:5822–5833, 2002) and the direct interpretation of dipolar couplings (DIDCs) (Tolman in J Am Chem Soc 124:12020–12030, 2002). Both methods have been incorporated into iterative schemes, namely the self-consistent RDC based MFA (SCRM) (Lakomek et al. in J Biomol NMR 41:139–155, 2008) and iterative DIDC (Yao et al. in J Phys Chem B 112:6045–6056, 2008), with the goal of removing the influence of structural noise in the MFA and DIDC formulations. Here, we report a new iterative procedure entitled Optimized RDC-based Iterative and Unified Model-free analysis (ORIUM). ORIUM unifies theoretical concepts developed in the MFA, SCRM, and DIDC methods to construct a computationally less demanding approach to determine these structural and dynamic parameters. In all schemes, dynamic averaging reduces the actual magnitude of the alignment tensors complicating the determination of the absolute values for the generalized order parameters. To readdress this scaling issue that has been previously investigated (Lakomek et al. in J Biomol NMR 41:139–155, 2008; Salmon et al. in Angew Chem Int Edit 48:4154–4157, 2009), a new method is presented using only RDC data to establish a lower bound on protein motion, bypassing the requirement of Lipari–Szabo order parameters. ORIUM and the new scaling procedure are applied to the proteins ubiquitin and the third immunoglobulin domain of protein G (GB3). Our results indicate good agreement with the SCRM and iterative DIDC approaches and signify the general applicability of ORIUM and the proposed scaling for the extraction of inter-nuclear vector structural and dynamic content.  相似文献   

2.
We prepared (2S,6S)-Z-Dpm(Z)(OMe) (4) by protease-mediated hydrolysis of (R,R/S,S)-Z-Dpm(Z)(OMe)-OMe (3), converted it to (2S,6S)-Dpm(Z)(OMe) (6) via PCI5 to an NCA intermediate and hydrolysis, protected the amino group with Boc to give (2S,6S)-Boc-Dpm(Z)(OMe) (7), which upon ammonolysis of the Me ester afforded (2S,6S)-Boc-Dpm(Z)(NH2) (8). Hydrogenolysis of 8 and protection with Fmoc gave (2S,6S)-Boc-Dpm(Fmoc)(NH2)(10). Using 10 and SPPS, we prepared three Dpm-containing peptides and their corresponding Lys peptides. Enzymatic studies with mLAP and cLAP showed that the Leu moiety in Ac-Gly-(2S,6S)-Dpm(Leu)(NH2)-Ala (14) was hydrolyzed 68-fold and >1000-fold more rapidly, respectively, than that in Ac-Gly-Lys(Leu)-Ala (12). The enhanced rate of Leu formation from 14 compared to 12 was also observed with homogenates of mouse C3 sarcomas. This homogenate also hydrolyzed Ac-Gly-(2S,6S)-Dpm(Ac-Gly-Pro-Gln-Gly-Leu)(NH2)-Ala (16) to Ac-Gly-(2S,6S)-Dpm(NH2)-Ala (13), Leu and Ac-Gly-Pro-Gln-Gly (17). This implies the side chain is cleaved first by endopeptidases, such as matrix metalloproteinases (MMPs), and then the remaining Leu is cleaved by LAP-like exopeptidases. The rate of liberation of 17 from 16 and the corresponding Lys isopeptide, Ac-Gly-Lys(Ac-Gly-Pro-Gln-Gly-Leu)-Ala (15), was not significantly different. The rate of formation of 13 was faster from 16 than Ac-Gly-Lys-Ala (11) was from 15. Thus, the entire isopeptide side chain can be removed by the cooperative action of LAP-like and MMP-like peptidases present in tumor tissue, which occurs faster in the Dpm peptide 16 than in the Lys peptide 15. The rate of formation of 13 from 16 by lung, liver, and intestine homogenates (from the same C3 tumor-bearing mice) was comparable to or higher than from the tumor homogenates, but the rate by blood was only 4% the value of the tumor homogenates. Analogs of a bioadhesive fragment from the laminin alpha1 chain were prepared by replacing the essential Lys with Dpm(NH2) (20) and Dpm(Leu)(NH2) (21). Both Dpm-containing peptides were active, although considerably weaker than the corresponding Lys peptides 18 and 19, in a cell attachment assay with human fibrosarcoma HT-1080 cells.  相似文献   

3.
Eleven new analogues were synthesized by modification of the potent oxytocin antagonist (OTA) [(S)Pmp(1), D-Trp(2), Pen(6), Arg(8)]-Oxytocin, or PA (parent antagonist), in which (S)Pmp = beta,beta-(3-thiapentamethylene)-beta-mercapto-propionic acid. By internal acylation of Lys, Orn, L-1,4-diaminobutyric acid (Dab), L-1,3-diaminopropionic acid (Dap) at position 4 with the C-terminal Gly of the peptide tail, we prepared cyclo-(4-9)-[Lys(4), Gly(9)]-PA (pA(2) = 8.77 +/- 0.27), 1, and cyclo-(4-9)-[Orn(4), Gly(9)]-PA (pA(2) = 8.81 +/- 0.25), 3, which are equipotent with PA (pA(2) = 8.68 +/- 0.18) in the rat uterotonic assay and cyclo-(4-9)-[Dab(4), Gly(9)]-PA, 4, cyclo-(4-9)-[Dap(4), Gly(9)]-PA, 5, and cyclo-(4-9)-[Pmp(1), Lys(4), Gly(9)]-PA, 2, which were weaker OTAs. Neither 1 nor 3 had activity as agonists or antagonists in the antidiuretic assay. In the pressor assay, both analogues 1 and 3, with pA(2) = 7.05 +/- 0.10 and pA(2) = 6.77 +/- 0.12, respectively, are somewhat weaker antagonists than PA (pA(2) = 7.47 +/- 0.35) showing significant gain in specificity. The [desamido(9)] PA-ethylenediamine monoamide, 6, and the dimer ([desamido(9)]-PA)(2) ethylenediamine diamide, 7, had lower potency in the uterotonic assay than PA. Additionally, we synthesized cyclo-(1-5)-[(HN)Pmp(1), Asp(5)]-PA, 8, inactive in all tests, which suggests that the intact Asn(5) side chain may be critical in the interaction of the OTAs with the oxytocin (OT) receptor. Similarly, cyclo-(5-9)-[Dap(5), Gly(9)]-PA, 9, had very low uterotonic potency. Two derivatives of PA truncated from the C-terminus were internally cyclized to Lys(4), giving rise to cyclo-(4-8)-desGly-NH(2)(9)[Lys(4), Arg(8)]-PA, 10 (pA(2) = 8.35 +/- 0.20), which maintains the high potency of PA and has no activity in the rat antidiuretic assay, and in the rat pressor assay it is about ten times weaker (pA2 = 6.41 +/- 0.15) than PA (pA2 = 7.47 +/- 0.35), thus showing gains in specificity, and to cyclo-(4-7)-desArg-Gly-(NH)(2)(8-9)[Lys(4), Pro(7))-PA, 11, which has much weaker potency than PA. Synthesis of cyclo-(4-6)-desPro-Arg-Gly-(NH)(2)(7-9)[Lys(4)]-PA failed.  相似文献   

4.
5.
6.
Emerging roles for Lys11-linked polyubiquitin in cellular regulation   总被引:1,自引:0,他引:1  
Polyubiquitin chains are assembled via one of seven lysine (Lys) residues or the N terminus. The cellular roles of Lys48- and Lys63-linked polyubiquitin have been extensively studied; however, the cellular functions of Lys11-linked chains are less well understood. Recent insights into Lys11-linked ubiquitin chains have revealed their important function in cell cycle control. Additionally, Lys11 linkages have been identified in the context of mixed chains in many other cellular pathways. In this review, we introduce the specific enzymes that mediate Lys11-linked chain assembly and disassembly, and discuss the diverse cellular processes in which Lys11 linkages participate. Notably, mechanistic insights have revealed how the E2 ubiquitin-conjugating enzyme UBE2S achieves its Lys11 linkage specificity, and two structures of Lys11-linked polyubiquitin highlight the dynamic nature of this compact chain type.  相似文献   

7.
Three isomeric complexes, trans-[PtCl2(NH3)(2-methylpiperidine)], trans-[PtCl2(NH3)(3-methylpiperidine)] and trans-[PtCl2(NH3)(4-methylpiperidine)], were prepared and their cytotoxicities against six ovarian cancer cell lines, three sensitive and three resistant to cisplatin, were measured. There were no significant differences in the cytotoxicities of the three isomers against these cell lines. The interactions of the three complexes with reduced glutathione (GSH) and with ubiquitin (Ub), as a model protein, were studied. The trans-[PtCl2(NH3)(2-methylpiperidine)] reacted approximately twice as slowly with GSH as did the other two isomers. In the 1:1 interactions of the three complexes with ubiquitin (Mr = 8565 amu), trans-[PtCl2(NH3)(3-methylpiperidine)] and trans-[PtCl2(NH3)(4-methylpiperidine)] attained 100% modification while trans-[PtCl2(NH3)(2-methylpiperidine)] reached only less than 50% modification. Trans-[PtCl2(NH3)(2-methylpiperidine)] reacts significantly less efficiently with GSH and proteins than the other two isomers yet this is not reflected in the cytotoxicity values. These results indicate that for these complexes, in these cell lines, cytosolic detoxification probably does not play a dominant role in determining the cytotoxicity of the complexes.  相似文献   

8.
Recruitment of substrates to the 26S proteasome usually requires covalent attachment of the Lys48‐linked polyubiquitin chain. In contrast, modifications with the Lys63‐linked polyubiquitin chain and/or monomeric ubiquitin are generally thought to function in proteasome‐independent cellular processes. Nevertheless, the ubiquitin chain‐type specificity for the proteasomal targeting is still poorly understood, especially in vivo. Using mass spectrometry, we found that Rsp5, a ubiquitin‐ligase in budding yeast, catalyzes the formation of Lys63‐linked ubiquitin chains in vitro. Interestingly, the 26S proteasome degraded well the Lys63‐linked ubiquitinated substrate in vitro. To examine whether Lys63‐linked ubiquitination serves in degradation in vivo, we investigated the ubiquitination of Mga2‐p120, a substrate of Rsp5. The polyubiquitinated p120 contained relatively high levels of Lys63‐linkages, and the Lys63‐linked chains were sufficient for the proteasome‐binding and subsequent p120‐processing. In addition, Lys63‐linked chains as well as Lys48‐linked chains were detected in the 26S proteasome‐bound polyubiquitinated proteins. These results raise the possibility that Lys63‐linked ubiquitin chain also serves as a targeting signal for the 26S proteaseome in vivo.  相似文献   

9.
10.
The anaphase-promoting complex or cyclosome (APC/C) initiates mitotic exit by ubiquitylating cell-cycle regulators such as cyclin B1 and securin. Lys 48-linked ubiquitin chains represent the canonical signal targeting proteins for degradation by the proteasome, but they are not required for the degradation of cyclin B1. Lys 11-linked ubiquitin chains have been implicated in degradation of APC/C substrates, but the Lys 11-chain-forming E2 UBE2S is not essential for mitotic exit, raising questions about the nature of the ubiquitin signal that targets APC/C substrates for degradation. Here we demonstrate that multiple monoubiquitylation of cyclin B1, catalysed by UBCH10 or UBC4/5, is sufficient to target cyclin B1 for destruction by the proteasome. When the number of ubiquitylatable lysines in cyclin B1 is restricted, Lys 11-linked ubiquitin polymers elaborated by UBE2S become increasingly important. We therefore explain how a substrate that contains multiple ubiquitin acceptor sites confers flexibility in the requirement for particular E2 enzymes in modulating the rate of ubiquitin-dependent proteolysis.  相似文献   

11.
The ubiquitin (Ub) fusion degradation pathway functions to degrade fusion proteins containing a nonremovable Ub moiety at their NH(2) terminus (Johnson, E. S., Ma, P. C., Ota, I. M., and Varshavsky, A. (1995) J. Biol. Chem. 270, 17442-17456). Here we show that ubiquitin fusion degradation also targets proteins for proteasomal degradation when Ub is present in the middle of fusion proteins (X-Ub-Y), in a process that entails polyubiquitylation of Ub Lys(48). By contrast, fusion proteins bearing COOH-terminal Ub (X-Ub) are metabolically stable. Such fusion proteins, either newly biosynthesized or generated by Ub hydrolases, are reversibly conjugated to heterogeneous target proteins in a manner similar to wild-type Ub. Most importantly, the NH(2)-terminal fusion partner (X) can maintain its structure and function in the formed X-Ub conjugates as inferred from the fluorescence of green fluorescent protein-Ub conjugates and the incorporation of human immunodeficiency virus type 1 Gag-Ub into viral particles. These findings strongly suggest that 26S proteasomes exhibit spatial discrimination of Ub-conjugated proteins, sparing domains extended from the NH(2) terminus of Ub from unfolding and degradation. The multifunctionality of X-Ub fusion proteins opens the possibility for a number of novel practical applications, including the imaging of Ub conjugate formation in living cells.  相似文献   

12.
The aquation and subsequent reactions of the dinuclear Pt antitumor complexes [{trans-PtCl(NH(3))(2)}(2)(mu-NH(2)(CH(2))(6)NH(2))](2+) (1,1/t,t) and [{cis-PtCl(NH(3))(2)}(2)(mu-NH(2)(CH(2))(6)NH(2))](2+) (1,1/c,c) in 15 mM perchlorate, acetate or phosphate solutions were followed at 298 K by [(1)H,(15)N] HSQC 2D NMR spectroscopy. Rate and equilibrium constants for the initial reversible aquation and the subsequent reversible reaction with phosphate or acetate are reported. The rate constant for the first aquation step is two-fold lower for 1,1/c,c than 1,1/t,t but the anation rate constants are similar so that the equilibrium lies further towards the chloro form for the 1,1/c,c compound. A pK (a) value of 6.01+/-0.03 was determined for the diaquated species [{cis-Pt(NH(3))(2)(H(2)O)}(2)(mu-NH(2)(CH(2))(6)NH(2))](4+) (1,1/c,c-3) which is 0.4 units higher than that of the 1,1/t,t compound. The rate constants for the binding of acetate and phosphate to 1,1/t,t are similar, but the rate constant for the reverse reaction is close to ten-fold higher in the case of phosphate so that equilibrium conditions are attained more rapidly (12 h compared with 64 h). On the other hand, for 1,1/c,c the rate constants for the forward and reverse reactions with acetate and phosphate are quite similar so that equilibrium conditions are reached very slowly (80-100 h) and a greater proportion of phosphate-bound species are present. The reduced lability of the bound phosphate for 1,1/c,c is attributed to the formation of a macrochelate phosphate-bridged species which was characterized by (31)P NMR and ESI-MS. The speciation profiles of 1,1/t,t and 1,1/c,c under physiological conditions are explored.  相似文献   

13.
Lys48-linked polyubiquitin chains serve as a signal for protein degradation by 26S proteasomes through its Ile44 hydrophobic patches interactions. The individual ubiquitin units of each chain are conjugated through an isopeptide bond between Lys48 and the C-terminal Gly76 of the preceding units. The conformation of Lys48-linked tetraubiquitin has been shown to change dynamically depending on solution pH. Here we enzymatically synthesized a wild-type Lys48-linked tetraubiquitin for structural study. In the synthesis, cyclic and non-cyclic species were obtained as major and minor fractions, respectively. This enabled us to solve the crystal structure of tetraubiquitin exclusively with native Lys48-linkages at 1.85 Å resolution in low pH 4.6. The crystallographic data clearly showed that the C-terminus of the first ubiquitin is conjugated to the Lys48 residue of the fourth ubiquitin. The overall structure is quite similar to the closed form of engineered tetraubiquitin at near-neutral pH 6.7, previously reported, in which the Ile44 hydrophobic patches face each other. The structure of the second and the third ubiquitin units [Ub(2)-Ub(3)] connected through a native isopeptide bond is significantly different from the conformations of the corresponding linkage of the engineered tetraubiquitins, whereas the structures of Ub(1)-Ub(2) and Ub(3)-Ub(4) isopeptide bonds are almost identical to those of the previously reported structures. From these observations, we suggest that the flexible nature of the isopeptide linkage thus observed contributes to the structural arrangements of ubiquitin chains exemplified by the pH-dependent closed-to-open conformational transition of tetraubiquitin.  相似文献   

14.
Ubiquitin chain complexity in cells is likely regulated by a diverse set of deubiquitinating enzymes (DUBs) with distinct ubiquitin chain preferences. Here we show that the polyglutamine disease protein, ataxin-3, binds and cleaves ubiquitin chains in a manner suggesting that it functions as a mixed linkage, chain-editing enzyme. Ataxin-3 cleaves ubiquitin chains through its amino-terminal Josephin domain and binds ubiquitin chains through a carboxyl-terminal cluster of ubiquitin interaction motifs neighboring the pathogenic polyglutamine tract. Ataxin-3 binds both Lys(48)- or Lys(63)-linked chains yet preferentially cleaves Lys(63) linkages. Ataxin-3 shows even greater activity toward mixed linkage polyubiquitin, cleaving Lys(63) linkages in chains that contain both Lys(48) and Lys(63) linkages. The ubiquitin interaction motifs regulate the specificity of this activity by restricting what can be cleaved by the protease domain, demonstrating that linkage specificity can be determined by elements outside the catalytic domain of a DUB. These findings establish ataxin-3 as a novel DUB that edits topologically complex chains.  相似文献   

15.
We report a new residual dipolar couplings (RDCs) based NMR procedure for rapidly determining RNA tertiary structure demonstrated on a uniformly (15)N/(13)C-labeled 27 nt variant of the trans-activation response element (TAR) RNA from HIV-I. In this procedure, the time-consuming nuclear Overhauser enhancement (NOE)-based sequential assignment step is replaced by a fully automated RDC-based assignment strategy. This approach involves examination of all allowed sequence-specific resonance assignment permutations for best-fit agreement between measured RDCs and coordinates for sub-structures in a target RNA. Using idealized A-form geometries to model Watson-Crick helices and coordinates from a previous X-ray structure to model a hairpin loop in TAR, the best-fit RDC assignment solutions are determined very rapidly (相似文献   

16.
Ubiquitin is a post-translational modifier that is involved in cellular functions through its covalent attachment to target proteins. Ubiquitin can also be conjugated to itself at seven lysine residues and at its amino terminus to form eight linkage-specific polyubiquitin chains for individual cellular processes. The Lys63-linked polyubiquitin chain is recognized by tandem ubiquitin-interacting motifs (tUIMs) of Rap80 for the regulation of DNA repair. To understand the recognition mechanism between the Lys63-linked diubiquitin (K63-Ub2) and the tUIMs in solution, we determined the solution structure of the K63-Ub2:tUIMs complex by using NOE restraints and RDC data derived from NMR spectroscopy. The structure showed that the tUIMs adopts a nearly straight and single continuous α-helix, and the two ubiquitin units of the K63-Ub2 separately bind to each UIM motif. The interfaces are formed between Ile44-centered patches of the two ubiquitin units and the hydrophobic residues of the tUIMs. We also showed that the linker region between the two UIM motifs possesses a random-coil conformation in the free state, but undergoes the coil-to-helix transition upon complex formation, which simultaneously fixes the relative position of ubiquitin subunits. These data suggest that the relative position of ubiquitin subunits in the K63-Ub2:tUIMs complex is essential for linkage-specific binding of Rap80 tUIMs.  相似文献   

17.
Chu HA  Feng YW  Wang CM  Chiang KA  Ke SC 《Biochemistry》2004,43(34):10877-10885
Light-induced Fourier transform infrared difference spectroscopy has been applied to studies of ammonia effects on the oxygen-evolving complex (OEC) of photosystem II (PSII). We found that NH(3) induced characteristic spectral changes in the region of the symmetric carboxylate stretching modes (1450-1300 cm(-1)) of the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectra of PSII. The S(2) state carboxylate mode at 1365 cm(-1) in the S(2)Q(A)(-)/S(1)Q(A) spectrum of the controlled samples was very likely upshifted to 1379 cm(-1) in that of NH(3)-treated samples; however, the frequency of the corresponding S(1) carboxylate mode at 1402 cm(-1) in the same spectrum was not significantly affected. These two carboxylate modes have been assigned to a Mn-ligating carboxylate whose coordination mode changes from bridging or chelating to unidentate ligation during the S(1) to S(2) transition [Noguchi, T., Ono, T., and Inoue, Y. (1995) Biochim. Biophys. Acta 1228, 189-200; Kimura, Y., and Ono, T.-A. (2001) Biochemistry 40, 14061-14068]. Therefore, our results show that NH(3) induced significant structural changes of the OEC in the S(2) state. In addition, our results also indicated that the NH(3)-induced spectral changes of the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are dependent on the temperature of the FTIR measurement. Among the temperatures we measured, the strongest effect was seen at 250 K, a lesser effect was seen at 225 K, and little or no effect was seen at 200 K. Furthermore, our results also showed that the NH(3) effects on the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are dependent on the concentrations of NH(4)Cl. The NH(3)-induced upshift of the 1365 cm(-1) mode is apparent at 5 mM NH(4)Cl and is completely saturated at 100 mM NH(4)Cl concentration. Finally, we found that CH(3)NH(2) has a small but clear effect on the spectral change of the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectrum of PSII. The effects of amines on the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectra (NH(3) > CH(3)NH(2) > AEPD and Tris) are inverse proportional to their size (Tris approximately AEPD > CH(3)NH(2) > NH(3)). Therefore, our results showed that the effects of amines on the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are sterically selective for small amines. On the basis of the correlations between the conditions (dependences on the excitation temperature and NH(3) concentration and the steric requirement for the amine effects) that give rise to the NH(3)-induced upshift of the 1365 cm(-)(1) mode in the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII and the conditions that give rise to the altered S(2) state multiline EPR signal, we propose that the NH(3)-induced upshift of the 1365 cm(-1) mode is caused by the binding of NH(3) to the site on the Mn cluster that gives rise to the altered S(2) state multiline EPR signal. In addition, we found no significant NH(3)-induced change in the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectrum at 200 K. Under this condition, the OEC gives rise to the NH(3)-stabilized g = 4.1 EPR signal and a suppressed g = 2 multiline EPR signal. Our results suggest that the structural difference of the OEC between the normal g = 2 multiline form and the NH(3)-stabilized g = 4.1 form is small.  相似文献   

18.
Oxidized flavodoxin from Cyanobacterium anabaena PCC 7119 is used as a model system to investigate the fast internal dynamics of a flavin-bearing protein. Virtually complete backbone and side chain resonance NMR assignments of an oxidized flavodoxin point mutant (C55A) have been determined. Backbone and side chain dynamics in flavodoxin (C55A) were investigated using (15)N amide and deuterium methyl NMR relaxation methods. The squared generalized order parameters (S(NH)(2)) for backbone amide N-H bonds are found to be uniformly high ( approximately 0.923 over 109 residues in regular secondary structure), indicating considerable restriction of motion in the backbone of the protein. In contrast, methyl-bearing side chains are considerably heterogeneous in their amplitude of motion, as indicated by obtained symmetry axis squared generalized order parameters (S(axis)(2)). However, in comparison to nonprosthetic group-bearing proteins studied with these NMR relaxation methods, the side chains of oxidized flavodoxin are unusually rigid.  相似文献   

19.
Ubiquitin plays essential roles in various cellular processes; therefore, it is of keen interest to study the structure-function relationship of ubiquitin itself. We investigated the modification of Lys(6) of ubiquitin and its physiological consequences. Mass spectrometry-based peptide mapping and N-terminal sequencing demonstrated that, of the 7 Lys residues in ubiquitin, Lys(6) was the most readily labeled with sulfosuccinimidobiotin. Lys(6)-biotinylated ubiquitin was incorporated into high molecular mass ubiquitin conjugates as efficiently as unmodified ubiquitin. However, Lys(6)-biotinylated ubiquitin inhibited ubiquitin-dependent proteolysis, as conjugates formed with Lys(6)-biotinylated ubiquitin were resistant to proteasomal degradation. Ubiquitins with a mutation of Lys(6) had similar phenotypes as Lys(6)-biotinylated ubiquitin. Lys(6) mutant ubiquitins (K6A, K6R, and K6W) also inhibited ATP-dependent proteolysis and caused accumulation of ubiquitin conjugates. Conjugates formed with K6W mutant ubiquitin were also resistant to proteasomal degradation. The dominant-negative effect of Lys(6)-modified ubiquitin was further demonstrated in intact cells. Overexpression of K6W mutant ubiquitin resulted in accumulation of intracellular ubiquitin conjugates, stabilization of typical substrates for ubiquitin-dependent proteolysis, and enhanced susceptibility to oxidative stress. Taken together, these results show that Lys(6)-modified ubiquitin is a potent and specific inhibitor of ubiquitin-mediated protein degradation.  相似文献   

20.
NMR spectroscopy was used to search for mechanistically significant differences in the local mobility of the main-chain amides of Bacillus circulans xylanase (BCX) in its native and catalytically competent covalent glycosyl-enzyme intermediate states. 15N T1, T2, and 15N[1H] NOE values were measured for approximately 120 out of 178 peptide groups in both the apo form of the protein and in BCX covalently modified at position Glu78 with a mechanism-based 2-deoxy-2-fluoro-beta-xylobioside inactivator. Employing the model-free formalism of Lipari and Szabo, the measured relaxation parameters were used to calculate a global correlation time (tau(m)) for the protein in each form (9.2 +/- 0.2 ns for apo-BCX; 9.8 +/- 0.3 ns for the modified protein), as well as individual order parameters for the main-chain NH bond vectors. Average values of the order parameters for the protein in the apo and complexed forms were S2 = 0.86 +/- 0.04 and S2 = 0.91 +/- 0.04, respectively. No correlation is observed between these order parameters and the secondary structure, solvent accessibility, or hydrogen bonding patterns of amides in either form of the protein. These results demonstrate that the backbone of BCX is well ordered in both states and that formation of the glycosyl-enzyme intermediate leads to little change, in any, in the dynamic properties of BCX on the time scales sampled by 15N-NMR relaxation measurements.  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