首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 125 毫秒
1.
Reaction of crotonaldehyde or two molecules of acetaldehyde with DNA generates 3-(2'-deoxyribos-1'-yl)-5,6,7,8-tetrahydro-8-hydroxy-6-methylpyrimido[1,2-a]purine-10(3H)one (2, Scheme 1), which occurs in (6R, 8R) and (6S, 8S) configurations (Fig. 1). These diastereomers were site-specifically incorporated into oligonucleotides, which were then inserted into a double-stranded DNA vector for genotoxicity studies. Modified DNA was introduced into human xeroderma pigmentosum A (XPA) cells to allow replication. Analysis of progeny plasmid revealed that these DNA adducts inhibit DNA synthesis to similar degrees. (6S, 8S)-2 miscodes more frequently than (6R, 8R)-2: 10% versus 5%. For both adducts, major miscoding events were G-->T transversions, but G-->A transitions were also observed at a comparable level for (6R, 8R)-2. G-->C transversions were the second most common events for (6S, 8S)-2. Comparison of these results with those of other 1,N2-propanodeoxyguanosine (PdG) adducts, which were evaluated by the same system, indicates that (i) their synthesis inhibiting potencies are stronger than that of the unsubstituted analog, 3-(2'-deoxyribos-1'-yl)-5,6,7,8-tetrahydro-8-hydroxypyrimido[1,2-a]purine-10(3H)one (1, Scheme 1), but weaker than that of 3-(2'-deoxyribos-1'-yl)-5,6,7,8-tetrahydro-6-hydroxypyrimido[1,2-a]purine-10(3H)one (3, Scheme 1); (ii) both isomers of 2 are more miscoding than 1; (iii) the miscoding potency of (6S, 8S)-2 is comparable to those of 3 and a model PdG 4 lacking a hydroxyl and a methyl group (Fig. 1). Therefore, considering the fact that 2 are formed endogenously as well as exogenously, they may play a significant role in aging and cancer in humans.  相似文献   

2.
Phenylacetic acid mustard (PAM; 2), a major metabolite of the anticancer agent chlorambucil (CLB; 1), was allowed to react with 2'-deoxyadenosine (dA), 2'-deoxyguanosine (dG), 2'-deoxycytidine (dC), 2'-deoxy-5-methylcytidine (dMeC), and thymidine (T) at physiological pH (cacodylic acid, 50% base). The reactions were followed by HPLC and analyzed by HPLC/MS and/or (1)H-NMR techniques. Although the predominant reaction observed was hydrolysis of PAM, 2 also reacted with various heteroatoms of the nucleosides to give a series of products: compounds 5-31. PAM (2) was found to be hydrolytically slightly more stable than CLB (1). The principal reaction sites of 2 with dA, dG, and with all pyrimidine nucleosides were N(1), N(7), and N(3), resp. Also, several other adducts were detected and characterized. There was no significant difference in the reactivity of 1 and 2 with dG, dA or T, but the N(3) dC-PAM adduct was deaminated easier than the corresponding CLB derivative. The role of PAM-2'-deoxyribonucleoside adducts on the cytotoxic and mutagenic properties of CLB (1) is discussed.  相似文献   

3.
Reaction of the rodent carcinogen acrylonitrile (AN) at pH 5.0 and/or pH 7.0 for 10 and/or 40 days with 2'-deoxyadenosine (dAdo), 2'-deoxycytidine (dCyd), 2'-deoxyguanosine (dGuo), 2'-deoxyinosine (dIno), N6-methyl-2'-deoxyadenosine (N6-Me-dAdo) and thymidine (dThd) resulted in the formation of cyanoethyl and carboxyethyl adducts. Adducts were not detected after 4 h. The adducts isolated were 1-(2-carboxyethyl)-dAdo (1-CE-dAdo), N6-CE-dAdo, 3-CE-dCyd, 7-(2-cyanoethyl)-Gua (7-CNE-Gua), 7,9-bis-CNE-Gua, imidazole ring-opened 7,9-bis-CNE-Gua, 1-CNE-dIno, 1-CE-N6-Me-dAdo and 3-CNE-dThd. Structures were assigned on the basis of UV spectra and electron impact (EI), chemical ionization (CI), desorption chemical ionization (DCI) and Californium-252 fission fragment ionization mass spectra. Evidence is presented which strongly suggests that N6-CE-dAdo was formed by Dimroth rearrangement of 1-CE-dAdo during the reaction between AN and dAdo. The carboxyethyl adducts resulted from initial cyanoethylation (by Michael addition) at a ring nitrogen adjacent to an exocyclic nitrogen atom followed by rapid hydrolysis of the nitrile moiety to a carboxylic acid. It was postulated that the facile hydrolysis is an autocatalyzed reaction resulting from the formation of a cyclic intermediate between nitrile carbon and exocyclic nitrogen. AN was reacted with calf thymus DNA (pH 7.0, 37 degrees C, 40 days) and the relative amounts of adducts isolated were 1-CE-Ade (26%), N6-CE-Ade (8%), 3-CE-Cyt (1%), 7-CNE-Gua (26%), 7,9-bis-CNE-Gua (4%), imidazole ring-opened 7,9-bis-CNE-Gua (19%) and 3-CNE-Thy (16%). Thus a carcinogen once adducted to a base in DNA was shown to be subsequently modified resulting in a mixed pattern of cyanoethylated and carboxyethylated AN-DNA adducts. Three of the adducts (1-CE-Ade, N6-CE-Ade and 3-CE-Cyt) were identical to adducts previously reported by us to be formed following in vitro reaction of the carcinogen beta-propiolactone (BPL) and calf thymus DNA. The results demonstrate that AN can directly alkylate DNA in vitro at a physiological pH and temperature.  相似文献   

4.
The linear noncompetitive inhibition of the pepsin-catalyzed hydrolysis of Ac-Phe-Phe-Gly at pH 2.1 by L-Ac-Phe, L-Ac-Phe-NH2, and L-Ac-Phe-OEt has been claimed to substantiate the ordered release of products specified by the amino-enzyme mechanism for pepsin action. According to this interpretation, the binding of inhibitor to free enzyme and the amino-enzyme intermediate (Scheme I) generates the observed inhibition pattern. The proposition is valid only if a simple alternative explanation for the kinetic data, Scheme II, can be disproved. Scheme II attributes the inhibition pattern to the binding of inhibitor to free enzyme and the enzyme-substrate (Michaelis) complex. The experiments reported here have enabled us to distinguish between the two mechanisms. The pepsin-catalyzed hydrolyses of Ac-Phe-Trp, Z-H'IS-Phe-Trp, Z-Gly-His-Phe-Trp, and Z-Ala-His-Phe-Trp at pH 1.8 occur exclusively at the Phe-Trp bond and must yield the same amino-enzyme, E-Trp, if it is implicated. Under these circumstances, Scheme I requires that a plot of 1/kc vs. (I)o for the four substrates and a given noncompetitive inhibitor provide a set of four parallel lines. Scheme II predicts that the four lines generally will not be parallel. L-Ac-Phe, L-Ac-Phe-NH2, L-Ac-Phe-OMe, and D-Ac-Phe act as linear noncompetitive inhibitors for the pepsin-catalyzed hydrolysis of the four Trp-containing substrates. The plot of 1/kc vs. (I)o for each inhibitor results in a set of four nonparallel lines. Therefore Scheme II must be correct and the detection of noncompetitive inhibition accompanying the pepsin-catalyzed hydrolysis of peptides offers no insight into the merits of the amino-enzyme hypothesis.  相似文献   

5.
The rates of the acid-catalyzed decarboxylation and amide hydrolysis of α-ketoglutaramic acid, the keto analog of glutamine, were investigated and the products of the reactions were characterized. In strong acid at 100°C, amide hydrolysis and decarboxylation occur with about equal facility, yielding α-ketoglutaric acid and 5-hydroxy-2-pyrrolidone, respectively. 5-Hydroxy-2-pyrrolidone undergoes further amide hydrolysis so that the products of complete acid hydrolysis of α-ketoglutaramic acid are ammonia (100%), carbon dioxide (50%), and approximately equal yields (50%) of α-ketoglutaric acid and succinic semialdehyde (β-formylpropionic acid). At increasing pH values, the relative rate of decarboxylation to amide hydrolysis of α-ketoglutaramic acid increases, such that, at pH values of 2 or greater, decarboxylation occurs almost exclusively. The decarboxylation product 5-hydroxy-2-pyrrolidone, was characterized chromatographically and by its infrared and pmr spectra; the compound may be regarded as the cyclized form of succinamic semialdehyde. A mechanism for the competing amide hydrolysis and decarboxylation reactions is proposed, and the potential biological significance of the decarboxylation pathway is discussed.  相似文献   

6.
The reaction between 2'-deoxycytidine and styrene 7,8-oxide (SO) resulted in alkylation at the 3-position and at the O(2)-position through the alpha- and beta-carbons of the epoxide but at the N(4)-position only through the alpha-carbon. The 3-alkylated adducts were found to deaminate to the corresponding 2'-deoxyuridine adducts (37 degrees C, pH 7.4) with half-lives of 6 min and 2.4 h for the alpha- and beta-isomers, respectively. The N(4)-alkylated products were stable at neutral pH. The O(2)-alkylated products were unstable being prone to depyrimidation and to isomerisation between alpha- and beta-isomers. In SO-treated double-stranded DNA, enzymatic hydrolysis allowed the identification of the beta3-deoxyuridine and alphaN(4)-deoxycytidine adducts (1.9 and 0.5% of total alkylation, respectively), in addition to the previously identified DNA-adducts. The 3-substituted uracil may have implications for the mutagenicity of SO.  相似文献   

7.
Previous investigations of the potential of metal-organic compounds as inhibitors of human immunodeficiency virus type I protease (HIV-1 PR) showed that the copper(II) complex diaqua [bis(2-pyridylcarbonyl)amido] copper(II) nitrate dihydrate and the complex bis[N2-(2,3,6-trimethoxybenzyl)-4-2-pyridinecarboxamide] copper(II) behaved as inhibitors of HIV-1 PR. In a search for similar readily accessible ligands, we synthesised and studied the structural properties of N2-(2-pyridylmethyl)-2-pyridinecarboxamide (L) copper(II) complexes. Three different crystal structures were obtained. Two were found to contain ligand L simultaneously in a tridentate and bidentate conformation [Cu(L(tri)L(bi))]. The other contained two symmetry-related ligands, coordinated through the pyridine nitrogen and the amide oxygen atoms [Cu(L(bi))(2)]. A search of the Cambridge Structural Database indicated that L(tri) resulting from nitrogen bound amide hydrogen metal substitution is favoured over chelation through the amide oxygen atom. In our case, we calculated that the conformation of L(tri) is 11 kcal/mol more favourable than that of L(bi). ESI-MS experiments showed that the Cu(L(bi))(2) structure could not be observed in solution, while Cu(L(tri)L(bi))-related complexes were indeed present. The lack of protease inhibition of the pyridine carboxamide copper(II) complexes was explained by the fact that the Cu(L(bi)L(tri)) complex could not fit into the HIV-1 active site.  相似文献   

8.
The rates of hydrolysis of the ester, amide and anilide substrates of p-guanidino-L-phenylalanine (GPA) by Streptomyces griseus trypsin (S. griseus trypsin) were compared with those of arginine (Arg) substrates. The specificity constant (kcat/km) for the hydrolysis of GPA substrates by the enzyme was 2-3-times lower than that for arginine substrates. The kcat and Km values for the hydrolysis of N alpha-benzoyl-p-guanidino-L-phenylalanine ethyl ester (Bz-GPA-OEt) by S. griseus trypsin are in the same order of magnitude as those of N alpha-benzoyl-L-arginine ethyl ester (Bz-Arg-OEt), although both values for the former when hydrolyzed by bovine trypsin are higher by one order of magnitude than those for the latter. The specificity constant for the hydrolysis of Bz-GPA-OEt by S. griseus trypsin is much higher than that for N alpha-benzoyl-p-guanidino-L-phenylglycine ethyl ester (Bz-GPG-OEt). As with the kinetic behavior of bovine trypsin, low values in Km and kcat were observed for the hydrolysis of amide and anilide substrates of GPA by S. griseus trypsin compared with those of arginine substrates. The rates of hydrolysis of GPA and arginine substrates by S. griseus trypsin are about 2- to 62-times higher than those obtained by bovine trypsin. Substrate activation was observed with S. griseus trypsin in the hydrolysis of Bz-GPA-OEt as well as Bz-Arg-OEt, whereas substrate inhibition was observed in three kinds of N alpha-protected anilide substrates of GPA and arginine. In contrast, no activation by the amide substrate of GPA could be detected with this enzyme.  相似文献   

9.
D Payet  F Gaucheron  M Sip    M Leng 《Nucleic acids research》1993,21(25):5846-5851
Single- and double-stranded oligonucleotides containing a single monofunctional cis-[Pt(NH3)2(dG)(N7-N-methyl-2-diazapyrenium)]3+ adduct have been studied at two NaCl concentrations. In 50 mM and 1 M NaCl, the adducts within the single-stranded oligonucleotides are stable. In contrast, they are unstable within the corresponding double-stranded oligonucleotides. In 50 mM NaCl, the bonds between platinum and guanine or N-methyl-2,7-diazapyrenium residues are cleaved and subsequently, intra- or interstrand cross-links are formed as in the reaction between DNA and cis-DDP. In 1 M NaCl, the main reaction is the replacement of N-methyl-2,7-diazapyrenium residues by chloride which generates double-stranded oligonucleotides containing a single monofunctional cis-[Pt(NH3)2(dG)Cl]+ adduct. The rates of closure of these monofunctional adducts to bifunctional cross-links have been studied in 60 mM NaClO4. Within d(TG.CT/AGCA), d(CG.CT/AGCG) and d(AG.CT/AGCT) (the symbol.indicates the location of the adducts in the central sequences of oligonucleotides), the half-lifes (t1/2) of the cis-[Pt(NH3)2(dG)Cl]+ adducts are respectively 12, 6 and 2.8 hr and the cross-linking reactions occur between guanine residues on the opposite strands. Within d(AG.TC/GACT), d(CG.AT/ATCG) and d(TGTG./CACA) or d(TG.TG/CACA) t1/2 are respectively 1.6, 8 and larger than 20 hr and the intrastrand cross-links are formed at the d(AG), d(GA) and d(GTG) sites, respectively. The conclusion is that the rates of conversion of cis-platinum-DNA monofunctional adducts to minor bifunctional cross-links are dependent on base sequence. The potential use of the instability of cis-[Pt(NH3)2(dG)(N7-N-methyl-2-diazapyrenium)]3+ adducts is discussed in the context of the antisense strategy.  相似文献   

10.
1:1 adducts of N,N′-bis(benzophenone)-1,2-diiminoethane (bz2en) with copper(I) chloride, bromide and iodide, [Cu(bz2en)2][CuX2] (X = Cl, Br, and I), have been synthesized and the structures of the solid bromide and iodide adducts were determined by X-ray crystallography from single-crystal data. The solid-state structure reveals ionic complexes containing a cation of copper(I) ion coordinated to four nitrogen atoms of two bz2en molecules (distorted tetrahedron) and a linear dibromocuprate(I) and a di-μ-iodo-diiododicuprate(I) anion for the bromo and iodo adducts, respectively. The bromo adduct structure contains CH?Br intermolecular hydrogen bonds. The complexes are very stable towards atmospheric oxygen in the solid state. The spectral properties of the above complexes are also discussed.  相似文献   

11.
B P Cho  L Zhou 《Biochemistry》1999,38(23):7572-7583
19F NMR spectroscopy was used to probe the conformation of a DNA adduct derived from the carcinogen 7-fluoro-N-acetyl-2-aminofluorene (FAAF) in three structural contexts: as a monomer and incorporated into single- and double-stranded DNA. The 19F NMR spectrum of dG-C8-FAAF [N-(deoxyguanosin-8-yl)-N-acetyl-7-fluoro-2-aminofluorene] in methanol at -30 degrees C exhibited four interconvertible signals in a 11:52:26:11 ratio. Dynamic NMR analysis indicated that the four torsional isomers arise from restricted rotation about the amide (gamma) (14.4 kcal/mol) and the guanyl-nitrogen (alpha) bonds. The conformational heterogeneity persisted in a single strand FAAF-12-mer, d(CTTCTTG[FAAF]ACCTC), whose 19F NMR spectrum at 22 degrees C and pH 7.0 gave only two signals in a 40:60 ratio, instead of four. The two 19F signals followed a two-site exchange with the rotation barrier of 14.7 kcal/mol about the amide (gamma') bond. A similar conformational theme was observed in the FAAF-12-mer duplex, d(CTTCTTG[FAAF]ACCTC).d(GAGGTCAAGAAG), which revealed two 19F resonances in a 41:59 ratio at 22 degrees C and pH 7.0. According to solvent-induced isotope and magnetic anisotropy effects, the two duplex conformers adopt exclusively a base displacement structure, being different only in their relative acetyl group orientations, cis (gamma' approximately 180 degrees) or trans (gamma' approximately 0 degrees ). Dynamic NMR data indicated that the two conformers do not exchange over a wide range of temperatures. This contrasts with the nonacetylated counterpart, which exhibits an equilibrium between the "B-type" and "stacked" conformers [Zhou, L., et al. (1997) J. Am. Chem. Soc. 119, 5384-5389]. The exclusive stacked nature of the AAF adducts may provide insight into why AAF adducts are more mutagenic and prone to repair than the nonacetylated AF adducts.  相似文献   

12.
The model ultimate carcinogen, trans-4-N-acetoxy-N-acetylaminostilbene (N-acetoxy-AAS), was reacted with guanosine (Guo) and deoxyguanosine (d-Guo) and the resulting adducts were purified by Sephadex LH-20 chromatography and HPLC for structure identification. A number of new adducts was identified by mass and 1H-NMR spectroscopy. The generation of all known adducts can now be explained by a common mechanism. The electrophile formed from the hydroxamic acid ester at C-beta reacts in a first step predominantly with N2 of guanine (Gua). The resulting quinone-imide intermediate reacts in a second step with either one of three nucleophiles: (1) predominantly with N3 of Gua to yield the previously described angular cyclic adducts ((5R,6R)/(5S,6S)-9-oxo-5,6,7,9-tetrahydro-imidazo(2,1-b)purines); (2) with N1 of Gua to yield linear cyclic adducts ((6R,7R)/(6S,7S)-9-oxo-5,6,7,9-tetrahydro-imidazo(1,2-a)purines); (3) with water to yield the open ring (1R,2R)/(1S,2S)-2-(N2'-guanyl)-1-hydroxyethanes. To some minor extent (1:8-1:9) the electrophile reacts first with N1 or N3 of guanine which leads to the formation of two pairs of the corresponding regioisomeric cyclic adducts. This reaction mechanism may also explain the formation of cross-links between different bases.  相似文献   

13.
alpha-, beta-, and gamma-cyclodextrins (CDs) completely inhibited raw starch digestion by glucoamylase I (GA I, MW 90,000) from Aspergillus awamori var. kawachi, and inhibited by 85% the raw starch adsorption of GA I at the CD concentrations of 1-5 mM. CDs at 1-5 mM did not inhibit gelatinized starch hydrolysis by GA I, but at the concentration of 50 mM, they inhibited such hydrolysis slightly. GA I was specifically adsorbed onto CD-Sepharose 6B, but glucoamylase I' (GA I', MW 73,000), which does not adsorb onto or digest raw starch, from the same strain was not adsorbed onto that gel. The adsorption of the glucoamylases onto raw starch and CD-Sepharose 6B was correlated to their digestion of raw starch. The hydrophobic adsorption of GA I onto CDs and raw starch occurred competitively at the Cp region, which is on the C-terminal side of Gp-I in the site for raw starch affinity of GA I, and inclusion complexes were formed.  相似文献   

14.
The model ultimate carcinogen trans-4-N-acetoxy-N-acetylaminostilbene was reacted with guanosine, deoxyguanosine, RNA and DNA using differently labeled reactants. The nucleoside as well as the deoxynucleoside yielded predominantly four cyclic guanine adducts: (S,S)- and (R,R)-guanine-N2,beta-N3,alpha-N-acetyl-aminobibenzyl and the regioisomers with the N2,alpha-N3,beta-attachment in a ratio of 9:9:1:1. The same adducts predominate in RNA and DNA which demonstrates that guanine reacts most avidly among the bases. The stability of the N-glycosidic bond is quite different between ribosides and deoxyribosides. Under neutral conditions, the riboside derivatives are stable, whereas deoxyribose is cleaved off rather readily. As a consequence DNA depurinizes to some extent during the in vitro reaction and during enzymatic digestion. On the other hand, N2,N3-attachment of the acetylaminostilbene moiety to guanine appears to impair the activity of nucleases for steric reasons. This could explain the incomplete enzymatic hydrolysis of modified nucleic acids. The results provide an important basis for further investigations to identify the nucleic acid adducts generated in vivo.  相似文献   

15.
REV1, a Y family DNA polymerase (pol), is involved in replicative bypass past DNA lesions, so-called translesion DNA synthesis. In addition to a structural role as a scaffold protein, REV1 has been proposed to play a catalytic role as a dCTP transferase in translesion DNA synthesis past abasic and guanine lesions in eukaryotes. To better understand the catalytic function of REV1 in guanine lesion bypass, purified recombinant human REV1 was studied with two series of guanine lesions, N(2)-alkylG adducts (in oligonucleotides) ranging in size from methyl (Me) to CH(2)(6-benzo[a]pyrenyl) (BP) and O(6)-alkylG adducts ranging from Me to 4-oxo-4-(3-pyridyl)butyl (Pob). REV1 readily produced 1-base incorporation opposite G and all G adducts except for O(6)-PobG, which caused almost complete blockage. Steady-state kinetic parameters (k(cat)/K(m)) were similar for insertion of dCTP opposite G and N(2)-G adducts but were severely reduced opposite the O(6)-G adducts. REV1 showed apparent pre-steady-state burst kinetics for dCTP incorporation only opposite N(2)-BPG and little, if any, opposite G, N(2)-benzyl (Bz)G, or O(6)-BzG. The maximal polymerization rate (k(pol) 0.9 s(-1)) opposite N(2)-BPG was almost the same as opposite G, with only slightly decreased binding affinity to dCTP (2.5-fold). REV1 bound N(2)-BPG-adducted DNA 3-fold more tightly than unmodified G-containing DNA. These results and the lack of an elemental effect ((S(p))-2'-deoxycytidine 5'-O-(1-thiotriphosphate)) suggest that the late steps after product formation (possibly product release) become rate-limiting in catalysis opposite N(2)-BPG. We conclude that human REV1, apparently the slowest Y family polymerase, is kinetically highly tolerant to N(2)-adduct at G but not to O(6)-adducts.  相似文献   

16.
The reactivity of 4-[13C]- and 5-[13C]-5-chloro-2-methylisothiazol-3-one (MCI) and 2-methylisothiazol-3-one (MI) towards a model peptide and glutathione was followed by 13C and 1H[13C] NMR spectroscopy. Both molecules were found to react with GSH but in addition MCI was found to react with histidine and lysine to form adducts of a different nature. Reaction with histidine led to stable substitution adducts through an addition-elimination reaction at position 5 while reaction with lysine led to the formation of open adducts of the thioamide or amide type.  相似文献   

17.
Nitrogen mustard (HN2) is a bifunctional alkylating agent which is thought to cause cytotoxicity by covalently binding to DNA. Most studies to date have looked at qualitatively determining the presence of DNA–HN2 adducts from reactions with native DNA. The adduct which is predominately formed in these reactions is N-[2-(hydroxyethyl)-N-(2-(7-guaninyl)ethyl]methylamine (N7G). A simple and sensitive reversed-phase high-performance liquid chromatographic (HPLC) method for the determination of N7G from DNA using ultraviolet detection is described. DNA samples having been exposed to HN2 treatment were hydrolyzed and preseparated from high-molecular-mass material by filtration using a molecular mass cut-off of 3000. The mobile phase consisted of methanol–26 mM ammonium formate, pH 6.5 (24:76, v/v). N7G, as well as the internal standard, methoxyphenol, were separated within 30 min. The recovery of N7G after hydrolysis of the DNA reaction product was quantitative and limits of detection and quantification of 10 and 20 ng/ml, respectively, were calculated. The method was validated in DNA–HN2 dose response experiments. The N7G reaction product appears to be the first reaction product formed at lower ratios of HN2/DNA but its production plateaus at higher ratios of HN2/DNA probably due to increased formation of hitherto unknown adducts. The method is simple and sensitive and for this reason, may be suited for the determination of DNA/HN2 reaction products.  相似文献   

18.
Numerous studies have suggested relationships between myeloperoxidase, inflammation, and atherosclerosis. MPO-derived reactive chlorinating species (RCS) attack membrane plasmalogens releasing alpha-chloro-fatty aldehydes (alpha-Cl-FALDs) including 2-chlorohexadecanal (2-ClHDA). The molecular targets of alpha-Cl-FALDs are not known. The current study demonstrates 2-ClHDA adducts with ethanolamine glycerophospholipids and Fmoc-lysine. Utilizing electrospray ionization mass spectrometry, chlorinated adducts were observed that are apparent Schiff base adducts. Reduction of these Schiff base adducts with sodium cyanoborohydride resulted in a novel, stable adduct produced by the elimination of HCl. NMR further confirmed this structure. 2-ClHDA adducts with ethanolamine glycerophospholipids were also substrates for phospholipase D (PLD). The hydrolysis products were derivatized to pentafluorobenzoyl esters, and further structurally confirmed by GC-MS. Multiple molecular species of 2-ClHDA-N-modified ethanolamine glycerophospholipids were observed in endothelial cells treated with 2-ClHDA. These results show novel Schiff base adducts of alpha-Cl-FALDs with primary amines, which may represent an important fate of alpha-Cl-FALDs.  相似文献   

19.
In this study, the microwave-assisted copper(I)-catalyzed 1,3-dipolar cycloaddition reaction was used to synthesize peptide triazole-based polymers from two novel peptide-based monomers: azido-phenylalanyl-alanyl-lysyl-propargyl amide (1) and azido-phenylalanyl-alanyl-glycolyl-lysyl-propargyl amide (2). The selected monomers have sites for enzymatic degradation as well as for chemical hydrolysis to render the resulting polymer biodegradable. Depending on the monomer concentration in DMF, the molecular mass of the polymers could be tailored between 4.5 and 13.9 kDa (corresponding with 33-100 amino acid residues per polymer chain). As anticipated, both polymers can be enzymatically degraded by trypsin and chymotrypsin, whereas the ester bond in the polymer of 2 undergoes chemical hydrolysis under physiological conditions, as was shown by a ninhydrin-based colorimetric assay and MALDI-TOF analysis. In conclusion, the microwave-assisted copper(I)-catalyzed 1,3-dipolar cycloaddition reaction is an effective tool for synthesizing biodegradable peptide polymers, and it opens up new approaches toward the synthesis of (novel) designed biomedical materials.  相似文献   

20.
Tay-Sachs disease is an inborn lysosomal disease characterized by excessive cerebral accumulation of GM2. The catabolism of GM2 to GM3 in man requires beta-hexosaminidase A (HexA) and a protein cofactor, the GM2 activator. Thus, Tay-Sachs disease can be caused by the deficiency of either HexA or the GM2 activator. The same cofactor found in mouse shares 74.1% amino acid identity (67% nucleotide identity) with the human counterpart. Between the two activators, the mouse GM2 activator can effectively stimulate the hydrolysis of both GM2 and asialo-GM2 (GA2) by HexA and, to a lesser extent, also stimulate HexB to hydrolyze GA2, whereas the human activator is ineffective in stimulating the hydrolysis of GA2 (Yuziuk, J. A., Bertoni, C., Beccari, T., Orlacchio, A., Wu, Y.-Y., Li, S.-C., and Li, Y.-T. (1998) J. Biol. Chem. 273, 66-72). To understand the role of these two activators in stimulating the hydrolyses of GM2 and GA2, we have constructed human/mouse chimeric GM2 activators and studied their specificities. We have identified a narrow region (Asn(106)-Tyr(114)) in the mouse cDNA sequence that might be responsible for stimulating the hydrolysis of GA2. Replacement of the corresponding site in the human sequence with the specific mouse sequence converted the ineffective human activator into an effective chimeric protein for stimulating the hydrolysis of GA2. This chimeric activator protein, like the mouse protein, is also able to stimulate the hydrolysis of GA2 by HexB. The mouse model of human type B Tay-Sachs disease recently engineered by the targeted disruption of the Hexa gene showed less severe clinical manifestation than found in human patients. This has been considered to be the result of the catabolism of GM2 via converting it to GA2 and further hydrolysis of GA2 to lactosylceramide by HexB with the assistance of mouse GM2 activator protein. The chimeric activator protein that bears the characteristics of the mouse GM2 activator may therefore be able to induce an alternative catabolic pathway for GM2 in human type B Tay-Sachs patients.  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-23

京公网安备 11010802026262号