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 optineurin gene, OPTN, is one of the causative genes of primary open-angle glaucoma. Although oligomerization of optineurin in cultured cells was previously observed by gel filtration analysis and blue native gel electrophoresis (BNE), little is known about the characteristics of optineurin oligomers. Here, we aimed to analyze the oligomeric state of optineurin and factors affecting oligomerization, such as environmental stimuli or mutations in OPTN. Using BNE or immunoprecipitation followed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE), we demonstrated that both endogenous and transfected optineurin exist as oligomers, rather than monomers, in NIH3T3 cells. We also applied an in situ proximity ligation assay to visualize the self-interaction of optineurin in fixed HeLaS3 cells and found that the optineurin oligomers were localized diffusely in the cytoplasm. Optineurin oligomers were usually detected as a single band of a size equal to that of the optineurin monomer upon SDS-PAGE, while an additional protein band of a larger size was observed when cells were treated with H2O2. We showed that larger protein complex is optineurin oligomers by immunoprecipitation and termed it covalent optineurin oligomers. In cells expressing OPTN bearing the most common glaucoma-associated mutation, E50K, covalent oligomers were formed even without H2O2 stimulation. Antioxidants inhibited the formation of E50K-induced covalent oligomers to various degrees. A series of truncated constructs of OPTN was used to reveal that covalent oligomers may be optineurin trimers and that the ubiquitin-binding domain is essential for formation of these trimers. Our results indicated that optineurin trimers may be the basic unit of these oligomers. The oligomeric state can be affected by many factors that induce covalent bonds, such as H2O2 or E50K, as demonstrated here; this provides novel insights into the pathogenicity of E50K. Furthermore, regulation of the oligomeric state should be studied in the future.  相似文献   

2.
Many c-type cytochromes (cyts) can form domain-swapped oligomers. The positively charged Hydrogenobacter thermophilus (HT) cytochrome (cyt) c552 forms domain-swapped oligomers during expression in the Escherichia coli (E. coli) expression system, but the factors influencing the oligomerization remain unrevealed. Here, we found that the dimer of the negatively charged Shewanella violacea (SV) cyt c5 exhibits a domain-swapped structure, in which the N-terminal helix is exchanged between protomers, similar to the structures of the HT cyt c552 and Pseudomonas aeruginosa (PA) cyt c551 domain-swapped dimers. Positively charged horse cyt c and HT cyt c552 domain swapped during expression in E. coli, whereas negatively charged PA cyt c551 and SV cyt c5 did not. Oligomers were formed during expression in E. coli for HT cyt c552 attached to either a co- or post-translational signal peptide for transportation through the cytoplasm membrane, but not for PA cyt c551 attached to either signal peptide. HT cyt c552 formed oligomers in E. coli in the presence and absence of rare codons. More oligomers were obtained from the in vitro folding of horse cyt c and HT cyt c552 by the addition of negatively charged liposomes during folding, whereas the amount of oligomers for the in vitro folding of PA cyt c551 and SV cyt c5 did not change significantly by the addition. These results indicate that the protein surface charge affects the oligomerization of c-type cyts in cells; positively charged c-type cyts assemble on a negatively charged membrane, inducing formation of domain-swapped oligomers during folding.  相似文献   

3.
Amyloid fibrils are associated with many neurodegenerative diseases. It was found that amyloidogenic oligomers, not mature fibrils, are neurotoxic agents related to these diseases. Molecular mechanisms of infectivity, pathways of aggregation, and molecular structure of these oligomers remain elusive. Here, we use all-atom molecular dynamics, molecular mechanics combined with solvation analysis by statistical-mechanical, three-dimensional molecular theory of solvation (also known as 3D-RISM-KH) in a new MM-3D-RISM-KH method to study conformational stability, and association thermodynamics of small wild-type Aβ17-42 oligomers with different protonation states of Glu22, as well the E22Q (Dutch) mutants. The association free energy of small β-sheet oligomers shows near-linear trend with the dimers being thermodynamically more stable relative to the larger constructs. The linear (within statistical uncertainty) dependence of the association free energy on complex size is a consequence of the unilateral stacking of monomers in the β-sheet oligomers. The charge reduction of the wild-type Aβ17-42 oligomers upon protonation of the solvent-exposed Glu22 at acidic conditions results in lowering the association free energy compared to the wild-type oligomers at neutral pH and the E22Q mutants. The neutralization of the peptides because of the E22Q mutation only marginally affects the association free energy, with the reduction of the direct electrostatic interactions mostly compensated by the unfavorable electrostatic solvation effects. For the wild-type oligomers at acidic conditions such compensation is not complete, and the electrostatic interactions, along with the gas-phase nonpolar energetic and the overall entropic effects, contribute to the lowering of the association free energy. The differences in the association thermodynamics between the wild-type Aβ17-42 oligomers at neutral pH and the Dutch mutants, on the one hand, and the Aβ17-42 oligomers with protonated Glu22, on the other, may be explained by destabilization of the inter- and intrapeptide salt bridges between Asp23 and Lys28. Peculiarities in the conformational stability and the association thermodynamics for the different models of the Aβ17-42 oligomers are rationalized based on the analysis of the local physical interactions and the microscopic solvation structure.  相似文献   

4.
Oligomers of bacteriopheophytin (BPh) and bacteriochlorophyll (BChl) were formed in mixed aqueous-organic solvent systems, and in aqueous micelles of the detergent lauryldimethylamine oxide (LDAO). Conditions were found that gave relatively homogeneous samples of oligomers and that allowed quantitative comparisons of the spectroscopic properties of the monomeric and oligomeric pigments. The formation of certain types of oligomers is accompanied by a large bathochromic shift of the long-wavelength (Qy) absorption band of the BChl or BPh, and by a substantial increase in its dipole strength (hyperchromism). The hyperchromism of the Qy band occurs at the expense of the Soret band, which loses dipole strength. The Qx band shifts slightly to shorter wavelengths and also loses dipole strength. The CD spectrum in the near-infra-red (Qy) region becomes markedly nonconservative. (The net rotational strength in the Qy region is positive.) This also occurs at the expense of the bands at shorter wavelengths, which gain a net negative rotational strength. The spectroscopic properties of the oligomers resemble those of some of the BChl-protein complexes found in photosynthetic bacteria. The oligomerization of BPh in LDAO micelles is linked to the formation of large, cylindrical micelles that contain on the order of 105 LDAO molecules. However, the spectral changes probably occur on the formation of small oligomers of BPh; they begin to be seen when the micelles contain about 10 molecules of BPh. The BPh oligomers formed in LDAO micelles fluoresce at 865 nm, but the fluorescence yield is decreased about 40-fold, relative to that of monomeric BPh. The fluorescence yield is insensitive to the BPh/LDAO molar ratio, suggesting that the oligomers formed under these conditions are predominantly dimers. When the oligomers are excited with a short flash of light, they are converted with a low quantum yield into a metastable form. This transformation probably involves alterations in the geometry of the oligomer, but not full dissociation.  相似文献   

5.
Aβ17-42 (so-called p3) amyloid is detected in vivo in the brains of individuals with Alzheimer's disease or Down's syndrome. We investigated the polymorphism of Aβ17-42 oligomers based on experimental data from steady-state NMR measurements, electron microscopy, two-dimensional hydrogen exchange, and mutational studies, using all-atom molecular-dynamics simulation with explicit solvent. We assessed the structural stability and the populations. Our results suggest that conformational differences in the U-turn of Aβ17-42 lead to polymorphism in β-sheet registration and retention of an ordered β-strand organization at the termini. Further, although the parallel Aβ17-42 oligomer organization is the most stable of the conformers investigated here, different antiparallel Aβ17-42 organizations are also stable and compete with the parallel architectures, presenting a polymorphic population. In this study we propose that 1), the U-turn conformation is the primary factor leading to polymorphism in the assembly of Aβ17-42 oligomers, and is also coupled to oligomer growth; and 2), both parallel Aβ17-42 oligomers and an assembly of Aβ17-42 oligomers that includes both parallel and antiparallel organizations contribute to amyloid fibril formation. Finally, since a U-turn motif generally appears in amyloids formed by full proteins or long fragments, and since to date these have been shown to exist only in parallel architectures, our results apply to a broad range of oligomers and fibrils.  相似文献   

6.
The conformations of a series of l-phenylalanine oligomers having the general formula BOC-(Phe)n-OMe (n = 1–9) were investigated by circular dichroism in a number of solvent systems. These studies indicate that in trifluoroethanol and in hexafluoroisopropanol these oligomers probably form β-associated conformations beginning at the hexamer.  相似文献   

7.
The link between the size of soluble amyloid β (Aβ) oligomers and their toxicity to rat cerebellar granule cells (CGC) was investigated. Variation in conditions during in vitro oligomerization of Aβ1-42 resulted in peptide assemblies with different particle size as measured by atomic force microscopy and confirmed by dynamic light scattering and fluorescence correlation spectroscopy. Small oligomers of Aβ1-42 with a mean particle z-height of 1-2 nm exhibited propensity to bind to phospholipid vesicles and they were the most toxic species that induced rapid neuronal necrosis at submicromolar concentrations whereas the bigger aggregates (z-height above 4-5 nm) did not bind vesicles and did not cause detectable neuronal death. A similar neurotoxic pattern was also observed in primary cultures of cortex neurons whereas Aβ1−42 oligomers, monomers and fibrils were non-toxic to glial cells in CGC cultures or macrophage J774 cells. However, both oligomeric forms of Aβ1-42 induced reduction of neuronal cell densities in the CGC cultures.  相似文献   

8.
Manipulation of Rubisco within higher plants is complicated by the different genomic locations of the large (L; rbcL) and small (S; RbcS) subunit genes. Although rbcL can be accurately modified by plastome transformation, directed genetic manipulation of the multiple nuclear-encoded RbcS genes is more challenging. Here we demonstrate the viability of linking the S and L subunits of tobacco (Nicotiana tabacum) Rubisco using a flexible 40-amino acid tether. By replacing the rbcL in tobacco plastids with an artificial gene coding for a S40L fusion peptide, we found that the fusions readily assemble into catalytic (S40L)8 and (S40L)16 oligomers that are devoid of unlinked S subunits. While there was little or no change in CO2/O2 specificity or carboxylation rate of the Rubisco oligomers, their Kms for CO2 and O2 were reduced 10% to 20% and 45%, respectively. In young maturing leaves of the plastome transformants (called ANtS40L), the S40L-Rubisco levels were approximately 20% that of wild-type controls despite turnover of the S40L-Rubisco oligomers being only slightly enhanced relative to wild type. The reduced Rubisco content in ANtS40L leaves is partly attributed to problems with folding and assembly of the S40L peptides in tobacco plastids that relegate approximately 30% to 50% of the S40L pool to the insoluble protein fraction. Leaf CO2-assimilation rates in ANtS40L at varying pCO2 corresponded with the kinetics and reduced content of the Rubisco oligomers. This fusion strategy provides a novel platform to begin simultaneously engineering Rubisco L and S subunits in tobacco plastids.  相似文献   

9.
The use of ribonuclease U2 in RNA sequence determination   总被引:24,自引:0,他引:24  
The catalog of oligomers produced by ribonuclease T1 digestion ofEscherichi coli 16S ribosomal RNA has been determined by a new method that involves the use of ribonuclease U2 fromUstilago sphaerogena. The sequences for the larger T1 oligomers (8 or more bases) determined in this way differ in more than 50 % of the cases from those reported previously (determined by other methods).  相似文献   

10.
Summary NMR spectral studies on the HCN oligomers suggest the presence of carboxamide and urea groupings. The release of CO2, H2O, HCN, CH3CN, HCONH2 and pyridine on pyrolysis is consistent with the presence of these groupings as well as carboxylic acid groups. No basic primary amine groupings could be detected with fluorescamine. Hydrazinolysis of the HCN oligomers releases 10% of the amino acids normally released by acid hydrolysis. The oligomers give a positive biuret test but this is not due to the presence of peptide bonds. There is no conclusive evidence for the presence of peptide bonds in the HCN oligomers. No diglycine was detected on partial hydrolysis of the HCN oligomers at pH 8.5 suggesting that HCN oligomers were not a source of prebiotic peptides.Chemical Evolution 38. For the previous papers see Ferris JP, Rao RV, Newton TA (1979). J Org Chem 44:4378–4381, 4381–4385; Ferris JP, Edelson EH, Mount NM, Sullivan AE (1979) J Mol Evol 13:317–330  相似文献   

11.
  • 1.1. The antibodies produced against the capsular poly-N-acetylneuraminic acid (poly-Neu5Ac) of E. coliK-92 (α 2-8-, α 2-9-linked) were 100-fold less sensitive than those obtained against E. coli K-235 capsular polysaccharide (CP) (α 2-8-linked) and recognized both kinds of polymers to a similar extent.
  • 2.2. The partial hydrolysis of each purified polysaccharide revealed that E. coli K-92 CP is more labile at acidic pH than the polymer α 2-8-linked of E. coli K-235.
  • 3.3. The antisera against CP from E. coli K-92 bound its own oligomers in which the number of Neu5Ac units was higher than three, whereas they only cross-reacted with the oligomers derived from E. coli K-235 containing a number of residues higher than 12.
  • 4.4. The antisera against E. coli K-235 CP that recognized α 2–8 oligomers with a number of Neu5Ac residues higher than 5, also reacted, although very weakly, with those containing α2–8 and α 2–9 linkages in which the carbon length was higher than (Neu5Ac)3.
  • 5.5. Both types of antibodies were also able to recognize the native antigens in living bacteria and could be employed for the recognition of the type of linkage presents in different sialylpolymers.
  相似文献   

12.
Amyloid β-protein (A β) assembles into oligomers that play a seminal role in Alzheimer’s disease (AD), a leading cause of dementia among the elderly. Despite undisputed importance of A β oligomers, their structure and the basis of their toxicity remain elusive. Previous experimental studies revealed that the [K16A] substitution strongly inhibits toxicity of the two predominant A β alloforms in the brain, A β 40 and A β 42, whereas the [K28A] substitution exerts only a moderate effect. Here, folding and oligomerization of [A16]A β 40, [A28]A β 40, [A16]A β 42, and [A28]A β 42 are examined by discrete molecular dynamics (DMD) combined with a four-bead implicit solvent force field, DMD4B-HYDRA, and compared to A β 40 and A β 42 oligomer formation. Our results show that both substitutions promote A β 40 and A β 42 oligomerization and that structural changes to oligomers are substitution- and alloform-specific. The [K28A] substitution increases solvent-accessible surface area of hydrophobic residues and the intrapeptide N-to-C terminal distance within oligomers more than the [K16A] substitution. The [K16A] substitution decreases the overall β-strand content, whereas the [K28A] substitution exerts only a modest change. Substitution-specific tertiary and quaternary structure changes indicate that the [K16A] substitution induces formation of more compact oligomers than the [K28A] substitution. If the structure-function paradigm applies to A β oligomers, then the observed substitution-specific structural changes in A β 40 and A β 42 oligomers are critical for understanding the structural basis of A β oligomer toxicity and correct identification of therapeutic targets against AD.  相似文献   

13.
Fragile X syndrome is caused by expansion of a d(CGG) trinucleotide repeat sequence in the 5′ untranslated region of the first exon of the FMR1 gene. Repeat expansion is thought to be instigated by formation of d(CGG)n secondary structures. Stable FMR1 d(CGG)n runs in normal individuals consist of 6–52 d(CGG) repeats that are interrupted every 9–11 triplets by a single d(AGG) trinucleotide. By contrast, individuals having fragile X syndrome premutation or full mutation present >54–200 or >200–2000 monotonous d(CGG) repeats, respectively. Here we show that the presence of interspersed d(AGG) triplets diminished in vitro formation of bimolecular tetrahelical structures of d(CGG)18 oligomers. Tetraplex structures formed by d(CGG)n oligomers containing d(AGG) interspersions had lower thermal stability. In addition, tetraplex structures of d(CGG)18 oligomers interspersed by d(AGG) triplets were unwound by human Werner syndrome DNA helicase at rates and to an extent that exceeded the unwinding of tetraplex form consisting of monotonous d(CGG)18. Diminished formation and stability of tetraplex structures of d(AGG)-containing FMR1 d(CGG)2–50 tracts might restrict their expansion in normal individuals.  相似文献   

14.
C57BL/6 (H-2 b) mice, when immunized with the sequential polymer (T-G-A-Gly)n and its low-molecular-weight oligomers, respond only to theα-helical oligomers with molecular weights of 9700 and above. Only the sameα-helical oligomers were able to inhibit the homologous antigen-antibody reaction. Random copolymers of GA, GT, and GAT10 did not inhibit. In vitro stimulation of peritoneal exudate lymphocyte (PETLES) cultures showed that the cellular response (T-cell) against (T-G-A-Gly)n is antigen-specific. In vitro antigen-induced stimulation of whole spleen or lymph node lymphocytes indicated that (T-G-A-Gly)n might also be a B-cell mitogen.  相似文献   

15.
Alzheimer’s disease (AD) is a well-known neurodegenerative disease. Deposition of β-amyloid protein (Aβ) oligomers plays a crucial role in the disease progression. Previous studies showed that toxicity induced by Aβ oligomers in cultured neurons and adult rat brain was partially mediated by activation of glutamatergic N-methyl-d-aspartate receptors (NMDAR). Additionally, memantine, a noncompetitive NMDAR antagonist, can significantly improve cognitive functions in some AD patients. However, little is currently known about the potential role of NMDAR antagonist on the regulation of P-MARCKS protein to Aβ1?42 oligomers induced neurotoxicity. The protective effect and mechanism of NMDAR antagonist on primary neurons exposed to Aβ1?42 oligomers were investigated in the study. We have defined that the Aβ1?42 treatment decreased cell viability and increased apoptosis. Moreover, Aβ1?42 oligomers exposure increased P-MARCKS and PIP2 expressions, while decreased SYP expression. However, NMDAR antagonist pretreatment ameliorates Aβ1?42 oligomers induced neuronal apoptosis and partially reverses the expression of P-MARCKS, PIP2 and SYP. In conclusion, NMDAR antagonist may ameliorate neurotoxicity induced by Aβ1?42 oligomers through reducing neuronal apoptosis and protecting synaptic plasticity in rat primary neurons. The mechanism involved may be mediated by the variation of protein P-MARCKS.  相似文献   

16.
Huntington disease is caused by expanded polyglutamine sequences in huntingtin, which procures its aggregation into intracellular inclusion bodies (IBs). Aggregate intermediates, such as soluble oligomers, are predicted to be toxic to cells, yet because of a lack of quantitative methods, the kinetics of aggregation in cells remains poorly understood. We used sedimentation velocity analysis to define and compare the heterogeneity and flux of purified huntingtin with huntingtin expressed in mammalian cells under non-denaturing conditions. Non-pathogenic huntingtin remained as hydrodynamically elongated monomers in vitro and in cells. Purified polyglutamine-expanded pathogenic huntingtin formed elongated monomers (2.4 S) that evolved into a heterogeneous aggregate population of increasing size over time (100–6,000 S). However, in cells, mutant huntingtin formed three major populations: monomers (2.3 S), oligomers (mode s20,w = 140 S) and IBs (mode s20,w = 320,000 S). Strikingly, the oligomers did not change in size heterogeneity or in their proportion of total huntingtin over 3 days despite continued monomer conversion to IBs, suggesting that oligomers are rate-limiting intermediates to IB formation. We also determined how a chaperone known to modulate huntingtin toxicity, Hsc70, influences in-cell huntingtin partitioning. Hsc70 decreased the pool of 140 S oligomers but increased the overall flux of monomers to IBs, suggesting that Hsc70 reduces toxicity by facilitating transfer of oligomers into IBs. Together, our data suggest that huntingtin aggregation is streamlined in cells and is consistent with the 140 S oligomers, which remain invariant over time, as a constant source of toxicity to cells irrespective of total load of insoluble aggregates.  相似文献   

17.
Alzheimer’s disease (AD) is a neurodegenerative disorder characterized pathologically by the abnormal deposition of extracellular amyloid-β (Aβ) oligomers. However, the nature and precise mechanism of the toxicity of Aβ oligomers are not clearly understood. Aβ oligomers have been previously shown to cause a major loss of EphB2, a member of the EphB family of receptor tyrosine kinases. To determine the effect of EphB2 on Aβ oligomer-induced neurotoxicity and the underlying molecular mechanisms, we examined the EphB2 gene in cultured hippocampal neurons. Using a cellular model of AD, Aβ1–42 oligomers were confirmed to induce neurotoxicity in a time-dependent manner and result in a major decrease of EphB2. EphB2 overexpression could prevent the neurotoxicity of hippocampal neurons from exposure to Aβ1–42 oligomers for 1 h. Further analysis revealed that EphB2 overexpression increased synaptic NR1 and NR2B expression in Aβ1–42 oligomer-treated neurons. Moreover, EphB2 overexpression prevented Aβ1–42 oligomer-induced downregulation of dephosphorylated p38 MAPK and phosphorylated CREB. Together, these results suggest that EphB2 is a factor which protects hippocampal neurons against the toxicity of Aβ1–42 oligomers, and we infer that the protection of EphB2 is achieved by increasing the synaptic NMDA receptor level and downstream p38 MAPK and CREB signaling in hippocampal neurons. This study provides new molecular insights into the neuroprotective effect of EphB2 and highlights its potential therapeutic role in the management of AD.  相似文献   

18.
Oligosaccharides of ([4)-β-d-ManpA-(→]n) with a degree of polymerisation (DP) of 5, 10 and 15 were generated by partial acid hydrolysis of alginate mannuronan. These were subsequently purified by a combination of size-exclusion chromatography and semi-preparative high-performance anion-exchange chromatography with pulsed amperometric detection. The purity of the isolated oligosaccharides was greater than 96%. With automated operation of the chromatography system, milligram quantities can be generated over a period of a few days. Thus, our methodology now offers some significant advantages over earlier, including our own, protocols focused on uronic acid oligomers, where the final products are either not as pure or more starting material is needed to generate an equivalent yield of product. Removal of ammonium ions in collected fractions after size-exclusion chromatography and prior to freeze-drying was found to be essential to prevent the formation of imines and subsequent Maillard browning products.  相似文献   

19.
Microbial Degradation of Natural Rubber Vulcanizates   总被引:4,自引:2,他引:2       下载免费PDF全文
An actinomycete, Nocardia sp. strain 835A, grows well on unvulcanized natural rubber and synthetic isoprene rubber, but not on other types of synthetic rubber. Not only unvulcanized but also various kinds of vulcanized natural rubber products were more or less utilized by the organism as the sole source of carbon and energy. The thin film from a latex glove was rapidly degraded, and the weight loss reached 75% after a 2-week cultivation period. Oligomers with molecular weights from 104 to 103 were accumulated during microbial growth on the latex glove. The partially purified oligomers were examined by infrared and 1H nuclear magnetic resonance and 13C nuclear magnetic resonance spectroscopy, and the spectra were those expected of cis-1, 4-polyisoprene with the structure, OHC—CH2—[—CH2—C(—CH3)=CH —CH2—]n—CH2—C(=O)— CH3, with average values of n of about 114 and 19 for the two oligomers.  相似文献   

20.
The coat protein of Tobacco Mosaic Virus is covalently labeled with a maleimide spin label at the single SH-group of the protein. Saturation transfer electron paramagnetic resonance spectroscopy, a technique that is sensitive to very slow molecular motion with rotational correlation times τc in the range 10?7 to 10?3 sec, shows the dissociation of large oligomers of spin labeled protein with τc~10?4 sec at pH 5.5 to smaller oligomers at higher pH.  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