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 0 毫秒
1.

Telomerase, responsible for telomere synthesis, is expressed in ~ 90% of human tumor cells but seldom in normal somatic cells. In this study, inhibition by carbocyclic oxetanocin G triphosphate (C.OXT-GTP) and its analogues was investigated in order to clarify the susceptibility of telomerase to various nucleotide analogues. C.OXT-GTP competitively inhibited telomerase activity with respect to dGTP. However, C.OXT-GTP had a potent inhibitory effect on DNA polymerase α. It was examined whether the nucleoside (C.OXT-G) was able to alter telomere length in cultured human HL60 cells. Contrary to expectation, long-term treatment with 10 μM C.OXT-G was found to cause telomere lengthening.  相似文献   

2.
Telomerase adds telomeric DNA repeats to the ends of linear chromosomal DNA. 3′-Azido-3′-deoxythymidine 5′-triphosphate (AZTTP) is a known telomerase inhibitor. To obtain more selective and potent inhibitors that can be employed as tools for studying telomerase, we investigated the telomerase-inhibitory effects of purine nucleosides bearing a 3′-down azido group: 3′-azido-2′,3′-dideoxyguanosine (AZddG) 5′-triphosphate (AZddGTP), 3′-azido-2′,3′-dideoxy-6-thioguanosine (AZddSG) 5′-triphosphate (AZddSGTP), 3′-azido-2′,3′-dideoxyadenosine (AZddA) 5′-triphosphate (AZddATP) and 3′-azido-2′,3′-dideoxy-2-aminoadenosine (AZddAA) 5′-triphosphate (AZddAATP). Of these, AZddGTP showed the most potent inhibitory activity against HeLa cell telomerase. AZddGTP was significantly incorporated into the 3′-terminus of DNA by partially purified telomerase. However, AZddGTP did not exhibit significant inhibitory activity against DNA polymerases α and δ, suggesting that AZddGTP is a selective inhibitor of telomerase.

We also investigated whether long-term treatment with these nucleosides could alter telomere length and growth rates of human HL60 cells in culture. Southern hybridization analysis of genomic DNA prepared from cells cultured in the presence of AZddG and AZddAA revealed reproducible telomere shortening.

  相似文献   

3.
The promyelocytic leukemia (HL60) cell line differentiates into monocyte-like cells after treatment with phorbol dibutyrate (PBt2). In contrast, bryostatin 1 (bryo), a structurally distinct protein kinase C (PKC) activator, does not induce differentiation and blocks the cytostatic effect of PBt2. The divergent responses to these agents correlate with activation of a PKC-like activity at the nucleus in response to bryo but not PBt2 (Fields, A. P., Pettit, G. R., and May, W.S. (1988) J. Biol. Chem. 263, 8253-8260). In the present study, this nuclear PKC-like activity (termed PKCn) was isolated from HL60 cells and shown to phosphorylate its known nuclear substrate, lamin B. PKCn-mediated phosphorylation of nuclear envelope-associated lamin B in vitro is calcium-dependent and is stimulated by bryo and 1,2-dioctanoylglycerol (DiC8), but not PBt2. In contrast, PKCn-mediated phosphorylation of histone IIIS is stimulated equally by all three activators. PKCn mediates calcium- and phosphatidylserine-dependent phosphorylation of both histone IIIS and partially purified lamin B. PKCn activity can be inhibited by an anti-PKC monoclonal antibody which specifically inhibits PKC. Isotype-specific PKC antibodies identify PKCn as beta II-PKC. Immunoblot analysis indicates that HL60 cells express both alpha- and beta II-PKC but no beta I- or gamma-PKC. Treatment of intact cells with bryo for 30 min leads to complete translocation of both alpha- and beta II-PKC from the cytosol to the membrane fractions. Approximately 8-10% of the total beta II-PKC (and less than 0.3% of the alpha-PKC) is found associated with the nuclear membrane of bryo-treated cells. In contrast, PBt2 treatment leads to complete translocation of alpha-PKC, but only partial translocation of beta II-PKC to the plasma membrane fraction. Neither PKC isotype is found associated with the nuclear membrane of PBt2-treated cells. These data demonstrate that alpha- and beta II-PKC differ with respect to activator responsiveness, intracellular distribution, and substrate specificity and indicate that their selective activation at distinct intracellular sites, including the nucleus, can have a dramatic effect on resulting cellular responses.  相似文献   

4.
Membrane interactions of tetradecapeptide toxin mastoparan (MP) and analogues (MP-3, MP-X and polistes MP), as indicated by inhibition of various enzymatic and cellular activities, were investigated. MP-3 was found to be the least active in inhibiting protein kinase C (PKC; activated by phosphatidylserine vesicles, synaptosomal membranes or phorbol ester), synaptosomal membrane Na,K-ATPase and proliferation and viability of leukemia HL60 cells. MP-3, however, was as active as others in inhibiting PKC activated by arachidonate monomers and phorbol ester binding. The unique properties of MP-3, the [des-Ile1-Asn2]-analogue of MP, might be related to its low functional amphiphilicity compared to others and useful in further delineating biological activities associated with or regulated by membranes.  相似文献   

5.
Inhibitors of protein phosphatases 1/2A (okadaic acid and calyculin A) exhibited differential cytotoxicity toward three human leukemia cell lines, in an increasing order of resistance, HL60 less than HL60/ADR less than K562 cells. Cytotoxicity of the toxins was associated with marked mitotic arrest of the cells, characterized by chromatid scattering/overcondensation and abnormal mitotic spindles. In all cases, calyculin A was more potent than okadaic acid. Protein phosphorylation experiments in intact cells revealed that HL60/ADR, the adriamycin-resistant variant, showed a higher overall phosphorylation of nuclear proteins than the drug-sensitive parental HL60, and that phorbol ester (protein kinase C activator) and calyculin A appeared to more specifically stimulate phosphorylation of p66 and p60, respectively. It was suggested that the toxins might be useful in delineating mechanisms underlying certain properties of cancer cells (such as multidrug resistance, mitosis and differentiation) related to protein phosphorylation/dephosphorylation reactions.  相似文献   

6.
Reactive oxygen species (ROS) including hydrogen peroxide (H(2)O(2)) are generated constitutively in mammalian cells. Because of its relatively long life and high permeability across membranes, H(2)O(2) is thought to be an important second messenger. Generation of H(2)O(2) is increased in response to external insults, including radiation. Catalase is located at the peroxisome and scavenges H(2)O(2). In this study, we investigated the role of catalase in cell growth using the H(2)O(2)-resistant variant HP100-1 of human promyelocytic HL60 cells. HP100-1 cells had an almost 10-fold higher activity of catalase than HL60 cells without differences in levels of glutathione peroxidase, manganese superoxide dismutase (MnSOD), and copper-zinc SOD (CuZnSOD). HP100-1 cells had higher proliferative activity than HL60 cells. Treatment with catalase or the introduction of catalase cDNA into HL60 cells stimulated cell growth. Exposure of HP100-1 cells to a catalase inhibitor resulted in suppression of cell growth with concomitant increased levels of intracellular H(2)O(2). Moreover, exogenously added H(2)O(2) or depletion of glutathione suppressed cell growth in HL60 cells. Extracellular signal regulated kinase 1/2 (ERK1/2) was constitutively phosphorylated in HP100-1 cells but not in HL60 cells. Inhibition of the ERK1/2 pathway suppressed the growth of HP100-1 cells, but inhibition of p38 mitogen-activated protein kinase (p38MAPK) did not affect growth. Moreover, inhibition of catalase blocked the phosphorylation of ERK1/2 but not of p38MAPK in HP100-1 cells. Thus our results suggest that catalase activates the growth of HL60 cells through dismutation of H(2)O(2), leading to activation of the ERK1/2 pathway; H(2)O(2) is an important regulator of growth in HL60 cells.  相似文献   

7.
The human leukemia cell line HL60 which resembles promyelocytes can be induced to differentiate to cells displaying features of the mature myeloid phenotype by a variety of agents including retinoic acid (RA) and agents that elevate intracellular adenosine 3:5 cyclic monophosphate (cyclic AMP) levels, e.g., 8-bromo-cyclic adenosine 3:5 monophosphate (8-Br-cyclic AMP), cholera toxin. Since most, if not all the effects of cyclic AMP, are mediated by adenosine 3:5 cyclic monophosphate-dependent protein kinase (cyclic AMP-dPK), we investigated the role of cyclic AMP-dPK and adenosine 3:5 cyclic monophosphate-independent protein kinase (cyclic AMP-iPK) in the induced differentiation of HL60 cells. Marked stimulation of cyclic AMP-dPK and cyclic AMP-iPK appears to be intimately involved with and specific for HL60 myeloid differentiation as evidenced by: (1) Stimulation of cyclic AMP-dPK and cyclic AMP-iPK early during HL60 myeloid differentiation and prior to phenotypic changes. (2) RA and dimethylformamide (DMF), agents that induce differentiation along the myeloid pathway, cause a marked increase in the type I cytosolic cyclic AMP-dPK and cyclic AMP-iPK (protamine kinase) while no such increases are noted in cells treated with 12-0-tetradecanoyl-phorbol-13-acetate (TPA) which induces differentiation along the monocyte/macrophage pathway. (3) Both native polyacrylamide gel electrophoresis as well as photoaffinity labeling with 8-azido-cyclic AMP demonstrate marked increases in type I cyclic AMP-dPK in the cytosols of cells exposed to agents that induce myeloid differentiation but no increase in TPA-differentiated cells. (4) The appearance and disappearance of specific cyclic AMP-dependent and -independent protein phosphorylations are associated with the induced myeloid differentiated state.  相似文献   

8.
M Tohkin  T Iiri  M Ui  T Katada 《FEBS letters》1989,255(1):187-190
Human promyelocytic leukemic (HL-60) cells were induced to differentiate into neutrophil- or macrophage-like cells by incubation of the cells with retinoic acid, dibutyryl cyclic AMP (Bt2cAMP) or phorbol 12-myristate 13-acetate (PMA). Differentiation was determined by an increase in the percentage of morphologically mature cells. The retinoic acid-induced differentiation of HL-60 cells was, but the Bt2cAMP- or PMA-induced one was not, inhibited by prior exposure of the cells to islet-activating protein (IAP), pertussis toxin. The IAP-induced inhibition was correlated with the toxin-catalyzed ADP-ribosylation of a membrane GTP-binding protein with a molecular mass of 40 kDa. Thus, the IAP-substrate GTP-binding protein appears to be involved in the retinoic acid-induced differentiation of HL-60 cells.  相似文献   

9.
The in vivo biological activity of four pharmaceutical preparations of recombinant human erythropoietin was compared. Two of the erythropoietins were produced by Chinese hamster ovary cells, CHO-K1, and the others were produced by mouse mammary cells, C127, and baby hamster kidney cells, BHK-21. The activities of the analogues were estimated by a simple cell counting method with conventional automated microcell counters. The amounts of these analogues gave straight logarithmic dose-response curves when plotted against the count of particles resistant to hemolysing reagent, which particles were mostly immature reticulocytes. The lines from the four analogues were parallel to each other. The relative activities of these analogues were 1.02, 1.19 and 1.21 when one of the analogues was arbitrarily used as the standard. These differences in the extent of the activity were not significant. Thus, the four recombinant human erythropoietin analogues, produced by four different mammalian cell lines, expressed the same biological potencies in vivo corresponding to their units, and the units used up to now by the manufacturers are equivalent. These results also draw the conclusion that the new simple in vivo bioassay can replace the existing accepted assay methods.  相似文献   

10.
Summary Human peripheral blood monocytes, upon activation, have the capacity to oxidize low density lipoprotein (LDL) and render the LDL toxic to cultured cells. Previous studies by our laboratory indicate that this process is mediated by free radicals in that it can be prevented by addition of free radical scavengers and antioxidants during the incubation of monocytes with LDL. Here we report that optimal modification of LDL by monocytes was influenced by media composition. In the absence of added metal ions, oxidation was distinctly dependent on the concentration of monocytes as well as LDL concentration. Exposure of monocytes to lipopolysaccharide or stimulation of phagocytosis by opsonized zymosan resulted in marked enhancement of LDL oxidation compared to other activating agents. After exposure to activated monocytes, lipid oxidation products in the supernatant were found both in a high molecular weight fraction containing LDL (>30 000 Daltons) and in a lipoprotein-free, low molecular weight fraction (<30 000 Daltons), yet only the high molecular weight, LDL-containing fraction was toxic to target cells. In addition, human myelomonocytic cell lines U937 and HL60 were shown to mediate oxidation of LDL. As with monocytes, exposing these cells to opsonized zymosan caused the level of LDL oxidation to be significantly enhanced. These findings offer further insight into the mechanisms of monocyte-mediated oxidation of lipoproteins and will facilitate studies investigating the role of monocyte-modified LDL in tissue injury. This project was funded by grants form the American Heart Association-Northeast Ohio Affiliate and the National Institutes of Health, Bethesda, MD (HL-29582).  相似文献   

11.
Differentiation therapy for neoplastic diseases has potential for supplementing existing treatment modalities but its implementation has been slow. One of the reasons is the lack of full understanding of the complexities of cellular pathways through which signals for differentiation lead to cell maturation. This was addressed in this study using HL60 cells, a well-established model of differentiation of neoplastic cells. SB 203580 and SB 202190, specific inhibitors of a signaling protein p38 MAP kinase, were found to markedly accelerate monocytic differentiation of HL60 cells induced by low concentrations of 1,25-dihydroxyvitamin D(3) (1,25D(3)). Surprisingly, inhibition of p38 activity resulted in sustained enhancement of p38 phosphorylation and of its in vitro activity in the absence of the inhibitor, indicating up-regulation of the upstream components of the p38 pathway. In addition, SB 203580 or SB 202190 treatment of HL60 cells resulted in a prolonged activation of the JNK and, to a lesser extent, the ERK pathways. The data are consistent with the hypothesis that in HL60 cells an interruption of a negative feedback loop from a p38 target activates a common regulator of multiple MAPK pathways. The possibility also exists that JNK and/or ERK pathways amplify a differentiation signal provided by 1,25D(3).  相似文献   

12.
During the lysis of leukemic cells with a monoclonal antibody cocktail (the so-called VIB pool) and complement the attempt was made to replace rabbit serum as a complement source by human serum. For identifying the lysis of leukemic cells the complement-dependent in vitro cytotoxicity test was used and for excluding stem cell toxicity the CFU-c test according to PIKE and ROBINSON. In combination with the applied monoclonal antibody pool against B and c-ALL the human complement could be shown to be suitable to produce a lysis in the same manner as rabbit complement. Similarly to the pretested rabbit serum the treatment with the human complement had no impact on stem cell recovery. An optimal cytotoxic activity (95% against ALL blasts of patients, 100% against NALM) could be identified up to an antibody dilution of 1:32 with a volume percentage of 50% of human complement, an incubation temperature of at least 37 degrees C and an incubation time of 30 mins. With proved high reactivity against leukemic cells and lacking impairment of the haemopoietic power of the bone-marrow, this method can be recommended for "purging" protocol with the possibility of using human serum as a source of complement having advantages as far as clinical application is concerned.  相似文献   

13.
We examined the ability of the synthetic selenium compound, 2-(4-methylphenyl)-1,3-selenazol-4-one (hereafter designated 3a), to induce apoptosis in a human ovarian cancer cell line (SKOV3) and a human leukemia cell line (HL-60). Flow cytometry showed that 3a treatment induced apoptosis in both cell lines to degrees comparable to that of the positive control, paclitaxel. Apoptosis was measured by PS externalization, DNA fragmentation and decreased mitochondrial membrane potential (MMP). However, analysis of the mechanism of action revealed differences between the responses of the two cell lines. Treatment with 3a arrested the cell cycle and induced caspase-3 activation in HL-60 cells, but not in SKOV3 cells. In contrast, 3a treatment induced apoptosis through translocation of AIF, a novel pro-apoptotic protein, in SKOV3 cells, but not in HL-60 cells. Collectively, our data demonstrated that 3a induced apoptosis in both cell lines, but via different action mechanisms.  相似文献   

14.
The relationship between phospholipase A2 and C activation and secretion was investigated in intact human neutrophils and differentiated HL60 cells. Activation by either ATP or fMetLeuPhe leads to [3H]arachidonic acid release into the external medium from prelabelled cells. This response was inhibited when the cells were pretreated with pertussis toxin. When the [3H]arachidonic acid-labelled cells were stimulated with fMetLeuPhe, ATP or Ca2+ ionophore A23187, and the lipids analysed by t.l.c., the increase in free fatty acid was accompanied by decreases in label from phosphatidylinositol and phosphatidylcholine. Moreover, incorporation of label into triacylglycerol and to a lesser extent phosphatidylethanolamine was evident. Activation of secretion was evident with ATP and fMetLeuPhe but not with A23187. The pharmacological specificity of the ATP receptor in HL60 cells was investigated by measuring secretion of beta-glucuronidase, formation of inositol phosphatases and release of [3H]arachidonic acid. External addition of ATP, UTP, ITP, adenosine 5'-[gamma-thio]triphosphate (ATP[S]), adenosine 5'-[beta gamma-imido]triphosphate (App[NH]p), XTP, CTP, GTP, 8-bromo-ATP and guanosine 5'-[gamma-thio]triphosphate (GTP[S]) to intact HL60 cells stimulated inositol phosphate production, but only the first five nucleotides were effective at stimulating secretion or [3H]arachidonic acid release. In human neutrophils, addition of ATP, ITP, UTP and ATP[S] also stimulated secretion from specific and azurophilic granules, and this was accompanied by increases in cytosolic Ca2+ and in [3H]arachidonic acid release. The addition of phorbol 12-myristate 13-acetate (PMA; 1 nM) prior to the addition of either fMetLeuPhe or ATP led to inhibition of phospholipase C activity. In contrast, this had no effect on phospholipase A2 activation, whilst secretion was potentiated. Phospholipase A2 activation by either agonist was dependent on an intact cell metabolism, as was secretion. It is concluded that (1) activation of phospholipase C does not always lead to activation of phospholipase A2, (2) phospholipase A2 is coupled to the receptor independently of phospholipase C via a pertussis-toxin-sensitive G-protein and (3) for secretion to take place, the receptor has to activate both phospholipases C and A2.  相似文献   

15.
The ability of human alpha and beta DNA polymerases and herpes simplex virus type 2 (HSV-2) and human cytomegalovirus (HCMV) DNA polymerases to insert and extend several nucleotide analogs has been investigated using a variation of Sanger-Coulson DNA sequencing technology. The analogs included the triphosphates of two antiviral nucleosides with incomplete sugar rings: 9-(1,3-dihydroxy-2-propoxymethyl)guanine (dhpG) and 9-(2-hydroxyethoxymethyl)guanine (acyG or acyclovir), as well as dideoxy and arabinosyl nucleoside triphosphates. Three pairs of contrasting behaviors were found, each pair distinguishing the two human polymerases from the two viral ones: first, extension behavior with araNTPs; second, insertion/extension behavior with dhpGTP; and third, the relative preference for insertion of ddGTP versus acyGTP. The relative level of insertion of the nucleotide analogs by HCMV and HSV-2 DNA polymerases was dhpGTP greater than (acyGTP and araNTP) greater than ddGTP, whereas by human alpha polymerase it was araATP greater than ddGTP much greater than (acyGTP and dhpGTP) and by human beta polymerase it was (araATP and ddGTP) much greater than (acyGTP and dhpGTP). Evidence is presented for three mechanisms of inhibition by extendible nucleotides (of dhp and ara types) exhibiting frequent internalization: araATP acted as a simple pseudoterminator of alpha and beta polymerases, but was easily extended past singlet sites by Herpesviridae polymerases and only stalled at sites requiring two or more araATP insertions in a row. Herpesviridae polymerases stalled after adding dhpGMP and one additional nucleotide, suggesting that polymerase translocation problems may be a factor in polymerase inhibition by modified sugar nucleotide analogs. The amino acid sequence of the human alpha DNA polymerase, which is acyGTP resistant, was found to vary by one amino acid from the amino sequences of the Herpesviridae polymerases in a region of significant similarity and probable functional homology. Amino acid differences at that same site differentiate acyclovir-resistant HSV-1 mutants from the acyclovir-sensitive HSV-1 wild type.  相似文献   

16.
The inhibition of highly purified herpes simplex virus (HSV)-induced and host cell DNA polymerases by the triphosphate form of 9-(2-hydroxyethoxymethyl)guanine (acyclovir; acycloguanosine) was examined. Acyclovir triphosphate (acyclo-GTP) competitively inhibited the incorporation of dGMP into DNA, catalyzed by HSV DNA polymerase; apparent Km and Ki values of dGTP and acyclo-GTP were 0.15 microM and 0.003 microM, respectively. HeLa DNA polymerase alpha was also competitively inhibited; Km and Ki values of dGTP and acyclo-GTP were 1.2 microM and 0.18 microM, respectively. In contrast, HeLa DNA polymerase beta was insensitive to the analogue. The "limited" DNA synthesis observed when dGTP was omitted from HSV or alpha DNA polymerase reactions was inhibited by acyclo-GTP in a concentration-dependent manner. Prior incubation of activated DNA, acyclo-GTP, and DNA polymerase (alpha or HSV resulted in a marked decrease in the utilization of the primer-template in subsequent DNA polymerase reactions. This decreased ability of preincubated primer-templates to support DNA synthesis was dependent on acyclo-GTP, enzyme concentration, and the time of prior incubation. Acyclo-GMP-terminated DNA was found to inhibit HSV DNA polymerase-catalyzed DNA synthesis. Kinetic experiments with variable concentrations of activated DNA and fixed concentrations of acyclo-GMP-terminated DNA revealed a noncompetitive inhibition of HSV-1 DNA polymerase. The apparent Km of 3'-hydroxyl termini was 1.1 X 10(-7) M, the Kii and Kis of acyclo-GMP termini in activated DNA were 8.8 X 10(-8) M and 2.1 X 10(-9) M, respectively. Finally, 14C-labeled acyclo-GMP residues incorporated into activated DNA by HSV-1 DNA polymerase could not be excised by the polymerase-associated 3',5'-exonuclease activity.  相似文献   

17.
Nucleoside analogues with modified sugar moieties have been examined for their substrate/inhibitor specificities towards highly purified deoxycytidine kinase (dCK) and thymidine kinases (tetrameric high-affinity form of TK1, and TK2) from human leukemic spleen. In particular, the analogues included the mono- and di-O'-methyl derivatives of dC, dU and dA, syntheses of which are described. In general, purine nucleosides with modified sugar rings were feebler substrates than the corresponding cytosine analogues. Sugar-modified analogues of dU were also relatively poor substrates of TK1 and TK2, but were reasonably good inhibitors, with generally lower Ki values vs TK2 than TK1. An excellent discriminator between TK1 and TK2 was 3'-hexanoylamino-2',3'-dideoxythymidine, with a Ki of approximately 600 microM for TK1 and approximately 0.1 microM for TK2. 3'-OMe-dC was a superior inhibitor of dCK to its 5'-O-methyl congener, consistent with possible participation of the oxygen of the (3')-OH or (3')-OMe as proton acceptor in hydrogen bonding with the enzyme. Surprisingly alpha-dT was a good substrate of both TK1 and TK2, with Ki values of 120 and 30 microM for TK1 and TK2, respectively; and a 3'-branched alpha-L-deoxycytidine analogue proved to be as good a substrate as its alpha-D-counterpart. Several 5'-substituted analogues of dC were good non-substrate inhibitors of dCK and, to a lesser extent, of TK2. Finally, some ribonucleosides are substrates of the foregoing enzymes; in particular C is a good substrate of dCK, and 2'-OMe-C is an even better substrate than dC.  相似文献   

18.
A series of non-steroidal anti-inflammatory drugs (NSAIDs) and some of their metabolites were evaluated for their ability to inhibit the proteolytic activity of human granulocyte elastase (HGE) and cathepsin G (HGC-G). The enzyme activity was monitored using specific synthetic chromogenic substrates. The results obtained indicated that phenylbutazone, sulindac, piroxicam and gold sodium thiomalate significantly inhibited HGE (Ki less than 0.5 mM), while only sulindac, diflunisal and gold sodium thiomalate were effective inhibitors of HGC-G (Ki less than 0.4 mM). Studies on metabolites of some of the NSAID tested were found to be superior inhibitors of both HGE and HGC-G than the parent molecules. Moreover, of the 18 compounds examined, the major metabolite of sulindac, sulindac sulphide was the most potent inhibitor of HGE (Ki = 0.01 mM) and HGC-G (Ki = 0.15 mM).  相似文献   

19.
We have previously reported on the assembly of recombinant human immunodeficiency virus (HIV)-like particles that contain gag structural proteins and present env glycoproteins gp120 and gp41 on their surfaces (O. Haffar,. J. Garriques, B. Travis, P. Moran, J. Zarling, and S.-L. Hu, J. Virol. 64:2653-2659, 1990). On the basis of their structures, we hypothesized that the recombinant particles would interfere with virus infection and tested our hypothesis in vitro by using peripheral blood mononuclear cells (PBMC) from HIV type 1-seropositive donors. Addition of the recombinant particles to PBMC concomitant with stimulation by anti-CD3 inhibited virus production, as determined by reduced levels of p24 in the culture supernatants. This inhibition of p24 production correlated with lower levels of cell-associated viral DNA. Several lines of evidence suggested that the recombinant particles exerted their antiviral effects primarily by inhibiting virus production from latently infected cells and not by inhibiting subsequent virus spread. Importantly, CD4+ T-cell stimulation by specific antigen or by anti-CD3 was not inhibited by treatment with the recombinant particles. This apparent selective inhibition of virus replication in infected PBMC represents a novel property of the recombinant HIV-like particles.  相似文献   

20.
The antipsychotic agent, remoxipride [(S)-(-)-3-bromo-N-[(1-ethyl-2-pyrrolidinyl)methyl]-2,6-dimethoxybenz amide] has been associated with acquired aplastic anemia. We have examined the ability of remoxipride, three pyrrolidine ring metabolites and five aromatic ring metabolites of the parent compound to induce apoptosis in HL60 cells and human bone marrow progenitor (HBMP) cells. Cells were treated for 0-24 h with each compound (0-200 microM). Apoptosis was assessed by fluorescence microscopy in Hoechst 33342- and propidium iodide stained cell samples. Results were confirmed by determination of internucleosomal DNA fragmentation using gel electrophoresis for HL60 cell samples and terminal deoxynucleotidyl transferase assay in HBMP cells. The catechol and hydroquinone metabolites, NCQ436 and NCQ344, induced apoptosis in HL60 and HBMP cells in a time- and concentration dependent manner, while the phenols, NCR181, FLA873, and FLA797, and the derivatives formed by oxidation of the pyrrolidine ring, FLA838, NCM001, and NCL118, had no effect. No necrosis was observed in cells treated with NCQ436 but NCQ344 had a biphasic effect in both cell types, inducing apoptosis at lower concentrations and necrosis at higher concentrations. These data show that the catechol and hydroquinone metabolites of remoxipride have direct toxic effects in HL60 and HBMP cells, leading to apoptosis, while the phenol metabolites were inactive. Similarly, benzene-derived catechol and hydroquinone, but not phenol, induce apoptosis in HBMP cells [Moran et al., Mol. Pharmacol., 50 (1996) 610-615]. We propose that remoxipride and benzene may induce aplastic anemia via production of similar reactive metabolites and that the ability of NCQ436 and NCQ344 to induce apoptosis in HBMP cells may contribute to the mechanism underlying acquired aplastic anemia that has been associated with remoxipride.  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