共查询到20条相似文献,搜索用时 15 毫秒
1.
Michio Tsuda Yukari Sei Masahiko Matsumoto Hiroshi Kamiguchi Masahiro Yamamoto Yukito Shinohara Tsuyoshi Igarashi Masaichi Yamamura 《Human genetics》1992,90(4):467-468
A new mutant -1-antichymotrypsin (variant ACT) was found by polymerase chain reaction single strand conformation polymorphism and direct sequencing. In this variant ACT, two bases (AA) were deleted from codon 391. This resulted in a different amino acid sequence downstream of the deletion point, elongating the peptide chain by 10 amino acids. 相似文献
2.
Emoto T Nakamura K Nagasaka Y Numa F Suminami Y Kato H 《Apoptosis : an international journal on programmed cell death》1998,3(3):155-160
Increased serum levels of α1-antichymotrypsin (α1ACT) are observed in some cancer patients, especially those with hepatocellular
carcinoma. A possible role of α1ACT in tumour growth has been suggested, but this remains uncertain. We have demonstrated
that α1ACT inhibited chymotrypsin-induced apoptosis in rat hepatoma H4 cells. Even low concentrations of chymotrypsin (but
not trypsin) induce apoptosis in H4 cells with a minimum effective concentration of 2.4 × 10−2 units/ml (0.5 μg/ml), and this
apoptosis was inhibited by α1ACT in a concentration-dependent manner. Furthermore, the concentrations of α1ACT required to
inhibit the apoptosis were lower than normal serum levels. These results may indicate that α1ACT plays a role in the apoptosis
of rat hepatoma cells.
This revised version was published online in June 2006 with corrections to the Cover Date. 相似文献
3.
Staphylococcal cysteine proteases are implicated as virulence factors in human and avian infections. Human strains of Staphylococcus aureus secrete two cysteine proteases (staphopains A and B), whereas avian strains express staphopain C (ScpA2), which is distinct from both human homologues. Here, we describe probable reasons why the horizontal transfer of a plasmid encoding staphopain C between avian and human strains has never been observed. The human plasma serine protease inhibitor α1-antichymotrypsin (ACHT) inhibits ScpA2. Together with the lack of ScpA2 inhibition by chicken plasma, these data may explain the exclusively avian occurrence of ScpA2. We also clarify the mechanistic details of this unusual cross-class inhibition. Analysis of mutated ACHT variants revealed that the cleavage of the Leu383-Ser384 peptide bond results in ScpA2 inhibition, whereas hydrolysis of the preceding peptide bond leads to ACHT inactivation. This evidence is consistent with the suicide-substrate-like mechanism of inhibition. 相似文献
4.
Yoshiyuki Sakano Mutsumi Sano Tsuneo Kobayashi 《Bioscience, biotechnology, and biochemistry》2013,77(12):3391-3398
Maltosyl-α-cyclodextrin (6-α-maltosylcyclomaltohexaose, M-CD) was prepared from maltose and α-cyclodextrin by the reverse action of Bacillus pullulanase, and the action of α-amylases on this dextrin was examined. Among α-amylases tested, Thermoactinomyces vulgaris α-amylase (TVA) and Taka-amylase A (TAA) were found to attack the M-CD. Their action pattern on M-CD was studied. These α-amylases cleaved, first the cyclodextrin ring of M-CD, and the branched octasaccharides formed were immediately degraded to form glucose, branched tetraose, or pentaose, though the action pattern was different for TVA and TAA. In addition, TAA also split M-CD into glucose and glucosyl-α-cyclodextrin. Fission products at various stages of the reaction were separated and analyzed by paper chromatography and high performance liquid chromatography, their structures were analyzed, and the degradation pattern of M-CD was found. 相似文献
5.
The differentiation of HL-60 cells induced by 1,25 dihydroxyvitamin D3 was found to be separated into two stages, i.e. commitment and promotion. Most of the HL-60 cells were committed to monocyte/macrophage lineage by pretreatment with 1,25 dihydroxyvitamin D3 (5–50 ng/ml) for 18–24 hr. The promotion in the second stage was inducer and lineage independent; treatment with 1.25% DMSO for 2 or 3 days promoted the differentiation of the committed HL-60 cells by 1,25 dihydroxyvitamin D3 into monocyte/macrophage lineage, but not granulocyte lineage.Abbreviations used NEA
nonspecific esterase activity
- NBT
nitroblue tetrazolium
- DMSO
dimethylsulfoxide
- RA
retinoic acid
- TPA
12-O-tetradecanoylphorbol-13-acetate 相似文献
6.
Background
The human pathogen Streptococcus pyogenes produces an endoglycosidase, EndoS that hydrolyzes the chitobiose core of the asparagine-linked glycan on the heavy chain of human IgG. IgG-binding to Fc gamma receptors (FcγR) on leukocytes triggers effector functions including phagocytosis, oxidative burst and the release of inflammatory mediators. The interactions between FcγR and the Fc domain of IgG depend on the IgG glycosylation state.Methodology/Principal Findings
Here we show for the first time that EndoS hydrolyzes the heavy chain glycan of all four human IgG subclasses (IgG1-4), in purified form and in a plasma environment. An inactive form of EndoS, obtained by site-directed mutagenesis, binds IgG with high affinity, in contrast to wild type EndoS that only transiently interacts with IgG, as shown by Slot-blotting and surface plasmon resonance technology. Furthermore, EndoS hydrolysis of the IgG glycan influences the binding of IgG to immobilized soluble FcγR and to an erythroleukemic cell line, K562, expressing FcγRIIa. Incubation of whole blood with EndoS results in a dramatic decrease of IgG binding to activated monocytes as analyzed by flow cytometry. Moreover, the IgG bound to K562 cells dissociates when cells are treated with EndoS. Likewise, IgG bound to immobilized FcγRIIa and subsequently treated with EndoS, dissociates from the receptor as analyzed by surface plasmon resonance and Western blot.Conclusions/Significance
We provide novel information about bacterial enzymatic modulation of the IgG/FcγR interaction that emphasizes the importance of glycosylation for antibody effector functions. Moreover, EndoS could be used as a biochemical tool for specific IgG N-glycan hydrolysis and IgG purification/detection, or as a potential immunosuppressing agent for treatment of antibody-mediated pathological processes. 相似文献7.
8.
Of 104 yeasts tested for the hydrolysis of rac-linalyl acetate and rac--terpinyl acetate twelve hydrolysed rac-linalyl acetate. Five of these belonged to five different species of Geotrichum. The most promising strain, Geotrichum capitatum CBS 0572.82, yielded (S)-linalool in 56% enantiomeric excess after 6h (E value 5, c = 32%). Only one unclassified yeast isolate hydrolyzed rac--terpinyl acetate with a slight preference for the S-enantiomer (E value 2). © Rapid Science Ltd. 1998 相似文献
9.
E. Bojarska R. Kraciuk J. Wierzchowski Z. Wieczorek J. Stępiński M. Jankowska 《Nucleosides, nucleotides & nucleic acids》2013,32(4-5):1125-1126
Abstract Hydrolysis of the following four cap analogs: m7G(5′)ppp(5′)A, m7G(5′)ppp(5′)m6A, m7G(5′)ppp(5′)m2′OG and m7G(5′)ppp(5′)2′dG catalyzed by homogeneous human Fhit protein and yellow lupin Ap3A hydrolase has been investigated. The hydrolysis products were identified by HPLC analysis and the Km and Vmax values calculated based on the data obtained by the fluorimetric method. 相似文献
10.
Jason E. Drury Luigi Di Costanzo Trevor M. Penning David W. Christianson 《The Journal of biological chemistry》2009,284(30):19786-19790
The Δ4-3-ketosteroid functionality is present in nearly all steroid hormones apart from estrogens. The first step in functionalization of the A-ring is mediated in humans by steroid 5α- or 5β-reductase. Finasteride is a mechanism-based inactivator of 5α-reductase type 2 with subnanomolar affinity and is widely used as a therapeutic for the treatment of benign prostatic hyperplasia. It is also used for androgen deprivation in hormone-de pend ent prostate carcinoma, and it has been examined as a chemopreventive agent in prostate cancer. The effect of finasteride on steroid 5β-reductase (AKR1D1) has not been previously reported. We show that finasteride competitively inhibits AKR1D1 with low micromolar affinity but does not act as a mechanism-based inactivator. The structure of the AKR1D1·NADP+·finasteride complex determined at 1.7 Å resolution shows that it is not possible for NADPH to reduce the Δ1-2-ene of finasteride because the cofactor and steroid are not proximal to each other. The C3-ketone of finasteride accepts hydrogen bonds from the catalytic residues Tyr-58 and Glu-120 in the active site of AKR1D1, providing an explanation for the competitive inhibition observed. This is the first reported structure of finasteride bound to an enzyme involved in steroid hormone metabolism.The Δ4-3-ketosteroid functionality is present in many important steroid hormones, e.g. testosterone, cortisone, and progesterone. An initial step in steroid hormone metabolism is the reduction of the Δ4-ene, which in humans is mediated by steroid 5α-reductases (SRD5A1, SRD5A2) or steroid 5β-reductase (AKR1D1)3 to yield the corresponding 5α- or 5β-dihydrosteroids, respectively (1, 2). The products of these reactions are not always inactive. 5α-Reductase is responsible for the conversion of testosterone to 5α-dihydrotestosterone (5α-DHT), which is the most potent natural ligand for the androgen receptor. By contrast, in addition to being involved in bile acid biosynthesis, 5β-reductase is responsible for generating 5β-pregnanes, which are natural ligands for the pregnane-X receptor (PXR) in the liver (3, 4). PXR is involved in the induction of CYP3A4, which is responsible for the metabolism of a large proportion of drugs (5, 6). Thus both 5α-reductase and 5β-reductase are involved in the formation of potent ligands for nuclear receptors.Finasteride is a selective 5α-reductase type 2 inhibitor that reduces plasma 5α-dihydrotestosterone levels and shrinks the size of the prostate (7). It is a widely used therapeutic agent in the treatment of benign prostatic hyperplasia (8, 9), it is used in androgen deprivation therapy to treat prostate cancer (10), and it has been examined as a chemopreventive agent for hormone-dependent prostate cancer (11). Finasteride was originally thought to act as a competitive inhibitor with nanomolar affinity for 5α-reductase type 2 (12). More recently, it was found that finasteride acts as a mechanism-based inactivator of this enzyme (13). Subsequent to inhibitor binding, there is hydride transfer from the NADPH cofactor to the Δ1-2-ene double bond of finasteride. The intermediate enolate tautomerizes at the enzyme active site to form a bisubstrate analogue in which dihydrofinasteride is covalently bound to NADP+ (13). The bisubstrate analogue has subnanomolar affinity for 5α-reductase type 2 (Fig. 1). No structural information exists for 5α-reductase type 1 or type 2; therefore, it is not possible to determine how finasteride would bind to the active site of a human steroid double bond reductase in the absence of an experimentally determined crystal structure.Open in a separate windowFIGURE 1.Mechanism-based inactivation of 5α-reductase type 2 by finasteride. Adapted from Bull et al. (13). R = −C(=O)-NH2; PADPR = 2′-phosphoadenosine-5″-diphosphoribose.Human steroid 5β-reductase is a member of the aldo-keto reductase (AKR) superfamily and is formally designated (AKR1D1) (14). The AKRs are soluble NADP(H)-dependent oxidoreductases with monomeric molecular masses of 37 kDa. These enzymes are amenable to x-ray crystallography, and during the last year, we and others have reported crystal structures of ternary complexes of AKR1D1 (15–17). The ternary complexes containing steroid substrates include: AKR1D1·NADP+·testosterone (PDB: 3BUR), AKR1D1·NADP+·progesterone (PDB: 3COT), AKR1D1·NADP+·cortisone (PDB: 3CMF), and AKR1D1·NADP+·Δ4-androstene-3,17-dione (PDB: 3CAS) (17). In addition, ternary complexes containing the products 5β-dihydroprogesterone (PDB: 3CAV) and 5β-dihydrotestosterone (PDB: 3DOP) have also been described (16, 18).As part of an ongoing inhibitor screen of AKR1D1, we now report that finasteride acts as a competitive inhibitor with low micromolar affinity. Additionally, we report the x-ray crystal structure of the AKR1D1·NADP+·finasteride complex. 相似文献
11.
Fusako Kawai Hideaki Yamada Koichi Ogata 《Bioscience, biotechnology, and biochemistry》2013,77(8):1963-1965
(—)-Epicatechin-3-gallate (ECG) and (— )-epigallocatechin-3-gallate (EGCG), major tea catechins, formed precipitates with soybean lipoxygenase (LOX) in the pH range of 4~7, although with accompanying 10 ~30% loss of the LOX activity. Yeast alcohol dehydrogenase also was precipitated by EGCG. Polyvinylpyrrolidone, Tween 20 and Triton X-100 dissociated the LOX activity from the EGCG-precipitated LOX. However, the MW of the dissociated LOX (114,000) differed from that of the native LOX (100,000). Enzyme activities of the EGCG-precipitated LOX and the dissociated LOX from the precipitate were less stable than the activity of the native LOX. These findings suggest the altered natures of proteins in the presence of tea catechins, ECG and EGCG. 相似文献
12.
Sequeira VB Rybchyn MS Tongkao-On W Gordon-Thomson C Malloy PJ Nemere I Norman AW Reeve VE Halliday GM Feldman D Mason RS 《Molecular endocrinology (Baltimore, Md.)》2012,26(4):574-582
UV radiation (UVR) is essential for formation of vitamin D(3), which can be hydroxylated locally in the skin to 1α,25-dihydroxyvitamin D(3) [1,25-(OH)(2)D(3)]. Recent studies implicate 1,25-(OH)(2)D(3) in reduction of UVR-induced DNA damage, particularly thymine dimers. There is evidence that photoprotection occurs through the steroid nongenomic pathway for 1,25-(OH)(2)D(3) action. In the current study, we tested the involvement of the classical vitamin D receptor (VDR) and the endoplasmic reticulum stress protein 57 (ERp57), in the mechanisms of photoprotection. The protective effects of 1,25-(OH)(2)D(3) against thymine dimers were abolished in fibroblasts from patients with hereditary vitamin D-resistant rickets that expressed no VDR protein, indicating that the VDR is essential for photoprotection. Photoprotection remained in hereditary vitamin D-resistant rickets fibroblasts expressing a VDR with a defective DNA-binding domain or a mutation in helix H1 of the classical ligand-binding domain, both defects resulting in a failure to mediate genomic responses, implicating nongenomic responses for photoprotection. Ab099, a neutralizing antibody to ERp57, and ERp57 small interfering RNA completely blocked protection against thymine dimers in normal fibroblasts. Co-IP studies showed that the VDR and ERp57 interact in nonnuclear extracts of fibroblasts. 1,25-(OH)(2)D(3) up-regulated expression of the tumor suppressor p53 in normal fibroblasts. This up-regulation of p53, however, was observed in all mutant fibroblasts, including those with no VDR, and with Ab099; therefore, VDR and ERp57 are not essential for p53 regulation. The data implicate the VDR and ERp57 as critical components for actions of 1,25-(OH)(2)D(3) against DNA damage, but the VDR does not require normal DNA binding or classical ligand binding to mediate photoprotection. 相似文献
13.
1. GM(1)-ganglioside, specifically tritiated in the terminal galactose, was hydrolysed by two forms of ;acid' methylumbelliferyl beta-galactosidase isolated on gel filtration. 2. Identification of GM(1)-ganglioside beta-galactosidase activity with the ;acid' methyl-umbelliferyl beta-galactosidases was based on the following: coincident elution profiles on gel filtration; simultaneous inactivation by heat and other treatments; stabilization of both activities by chloride ions; mutual inhibition of hydrolysis by the two substrates. 3. The two isoenzymes (I) and (II) showed general requirements for a mixture of anionic and nonionic detergents in the hydrolysis of the natural substrate. 4. Isoenzyme (I) differed from (II) in molecular size, pH-activity profile, relative resistance to dilution and in sensitivity to various inhibitors. 5. The most significant difference between the isoenzymes is in substrate saturation kinetics: (I) was hyperbolic whereas (II) was sigmoid. The apparent Michaelis constants were 28mum for (I) and 77mum for (II). Isoenzyme (I) was insensitive to GM(2)-ganglioside whereas (II) was inhibited, consistent with the hypothesis that GM(1)-ganglioside (and its analogue) acts as modifier in isoenzyme (II) but not in (I). 6. Isoenzyme (I) was membrane-bound whereas (II) was soluble; the former probably represents isoenzyme (II) bound to membrane components, thereby becoming activated. 7. Membranes may serve a dual role in enzyme catalysis involving lipids: as a medium where both enzyme and substrate are effectively concentrated, and as actual activator of enzymes through binding of the latter to specific membrane components. 相似文献
14.
T cells were isolated from peripheral blood lymphocytes (PBL) and sensitized to allogeneic PBL in a one-way mixed lymphocyte culture. The sensitized cells were fractionated on the basis of the presence of Fc receptors for IgG (TG+) or IgM (TM+), or the absence of both IgG and IgM receptors (TG?,M?). The Cytotoxicity of the T cells was found to reside principally in the TG?,M? subset. The degree of target cell lysis by this subset was related to the effector-to-target cell ratio. The sensitized T cells were also separated into subsets by treatment with monoclonal OKT4 antibody and complement (yielding OKT8+ cells) or OKT8 antibody and complement (yielding OKT4+ cells). The OKT8+ subset was the more cytotoxic of the two subsets, and this Cytotoxicity was again related to the effector-to-target cell ratio. The T-cell subsets obtained by these methods were characterized by immunologie and morphologic means. The Cytotoxicity of the total sensitized T-cell population or the TG?,M? subset could be neutralized to a considerable extent by anti-human α-lymphotoxin (anti-α-LT) serum, and α-LT thus appears to have an important role in cytolysis in this system. 相似文献
15.
This review analyzes data on the biological role of 3-hydroxysteroid dehydrogenase (3-HSD) in animal and human tissues and describes its main characteristics, mechanism of action, and regulation of activity. Based on published data, a scheme for the actions of androgen, progestin, and glucocorticoids involving the participation of 3-HSD is proposed. According to this scheme, in the mechanism of steroid action 3-HSD not only regulates the concentration of the main effector androgen, 5-dihydrotestosterone, in target cells, but also switches androgen, progestin, and glucocorticosteroid genomic activity to non-genomic activity. 相似文献
16.
Insulin-degrading enzyme (IDE) is involved in the clearance of many bioactive peptide substrates, including insulin and amyloid-β, peptides vital to the development of diabetes and Alzheimer's disease, respectively. IDE can also rapidly degrade hormones that are held together by intramolecular disulfide bond(s) without their reduction. Furthermore, IDE exhibits a remarkable ability to preferentially degrade structurally similar peptides such as the selective degradation of insulin-like growth factor (IGF)-II and transforming growth factor-α (TGF-α) over IGF-I and epidermal growth factor, respectively. Here, we used high-accuracy mass spectrometry to identify the cleavage sites of human IGF-II, TGF-α, amylin, reduced amylin, and amyloid-β by human IDE. We also determined the structures of human IDE-IGF-II and IDE-TGF-α at 2.3 Å and IDE-amylin at 2.9 Å. We found that IDE cleaves its substrates at multiple sites in a biased stochastic manner. Furthermore, the presence of a disulfide bond in amylin allows IDE to cut at an additional site in the middle of the peptide (amino acids 18-19). Our amylin-bound IDE structure offers insight into how the structural constraint from a disulfide bond in amylin can alter IDE cleavage sites. Together with NMR structures of amylin and the IGF and epidermal growth factor families, our work also reveals the structural basis of how the high dipole moment of substrates complements the charge distribution of the IDE catalytic chamber for the substrate selectivity. In addition, we show how the ability of substrates to properly anchor their N-terminus to the exosite of IDE and undergo a conformational switch upon binding to the catalytic chamber of IDE can also contribute to the selective degradation of structurally related growth factors. 相似文献
17.
18.
Hiromichi Takahashi Yoshihiro Hatta Noriyoshi Iriyama Yuichiro Hasegawa Hikaru Uchida Masaru Nakagawa Makoto Makishima Jin Takeuchi Masami Takei 《PloS one》2014,9(11)
Retinoids and 1α,25-dihydroxyvitamin D3 (1,25(OH)2D3) induce differentiation of myeloid leukemia cells into granulocyte and macrophage lineages, respectively. All-trans retinoic acid (ATRA), which is effective in the treatment of acute promyelocytic leukemia, can induce differentiation of other types of myeloid leukemia cells, and combined treatment with retinoid and 1,25(OH)2D3 effectively enhances the differentiation of leukemia cells into macrophage-like cells. Recent work has classified macrophages into M1 and M2 types. In this study, we investigated the effect of combined treatment with retinoid and 1,25(OH)2D3 on differentiation of myeloid leukemia THP-1 and HL60 cells. 9-cis Retinoic acid (9cRA) plus 1,25(OH)2D3 inhibited proliferation of THP-1 and HL60 cells and increased myeloid differentiation markers including nitroblue tetrazolium reducing activity and expression of CD14 and CD11b. ATRA and the synthetic retinoic acid receptor agonist Am80 exhibited similar effects in combination with 1,25(OH)2D3 but less effectively than 9cRA, while the retinoid X receptor agonist HX630 was not effective. 9cRA plus 1,25(OH)2D3 effectively increased expression of M2 macrophage marker genes, such as CD163, ARG1 and IL10, increased surface CD163 expression, and induced interleukin-10 secretion in myeloid leukemia cells, while 9cRA alone had weaker effects on these phenotypes and 1,25(OH)2D3 was not effective. Taken together, our results demonstrate selective induction of M2 macrophage markers in human myeloid leukemia cells by combined treatment with 9cRA and 1,25(OH)2D3. 相似文献
19.
1. Normal human serum was found to inhibit human cathepsin B1. 2. The major inhibitor present in serum was purified and identified as alpha(2)-macroglobulin. 3. alpha(2)-Macroglobulin was found to bind cathepsin B1 in an approximately 1:1 molar ratio. When bound, the enzyme retained about 50% of its proteolytic activity, and up to 80% of its activity against alpha-N-benzoyl-dl-arginine 2-naphthylamide. 4. Pretreatment of alpha(2)-macroglobulin with cathepsin B1 inactivated by exposure to pH8.5 or iodoacetic acid, in large molar excess, did not prevent the subsequent binding of active enzyme. Active enzyme, once bound, was not protected from inhibition by 1-chloro-4-phenyl-3-tosylamido-l-butan-2-one. 5. Cathepsin B1 was also inhibited by human immunoglobulin G, at high concentration. 6. Because it had been suggested that haptoglobin is responsible for the inhibition of ;cathepsin B' by serum, a method was devised for the selective removal of haptoglobin from mixtures of serum proteins by adsorption on haemoglobin covalently linked to Sepharose. No evidence was obtained that haptoglobin has any inhibitory activity against the enzyme. 相似文献
20.
Kyoung‐Soo Choi Liguo Song James Marshall Anders L. Lund Henry Shion 《Preparative biochemistry & biotechnology》2013,43(1):3-17
Abstract We compared the 2DE coupled to MALDI‐TOF‐MS and ESI‐MS/MS analysis (2DE‐MS) and the on‐line 2D nanoLC, followed by nanoESI‐MS/MS analysis (2DLC‐MS), for the separation and identification of proteins in high abundance protein‐depleted human plasma. Identification of proteins in the plasma by the two methods demonstrated that the majority of the identified protein set was unique to each method. Therefore, if a comprehensive coverage of the proteome identification is desired, it is ideal to apply both methods. The 2DE‐MS method is amenable to protein spot‐based quantitation, whereas the 2DLC‐MS method may provide an advantage of the high throughput application. 相似文献