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.
Mutants of Chinese hamster ovary cells were selected for resistance to a 3 hour exposure to 4 microgram/ml N-methyl-N'-nitro-N-nitrosoguanidine and tested for glutathione (GSH) levels. Six of eight clones that survived the initial treatment had reduced GSH levels ranging from 26 to 61% of wild-type values. These eight cell lines were tested for their susceptibility to a drug conjugate in which methotrexate (MTX) is disulfide-linked to poly(D-lysine) (MTX-SS-PDL) to test their capacity to cleave the endocytosed disulfide bond and release free MTX from this otherwise undegradable drug conjugate. We had shown that wild-type cells were killed by approximately 1 x 10(-7) M MTX given as free drug, as MTX-poly(L-lysine) or as MTX-SS-PDL, but were not affected by MTX-poly(D-lysine). All six lines with abnormally low levels of GSH were resistant to MTX-SS-PDL. The variants with the lowest levels of GSH (MNR-5 and MNR-10) were tested further and showed near-normal sensitivity to MTX and MTX poly(L-lysine). As expected, both lines were hypersensitive to melphalan. They were, however, normally sensitive to diphtheria toxin and ricin, indicating that some cleavage of the interchain disulfides in these protein toxins occurs even when cellular GSH is abnormally low. The lesser GSH requirement for toxin activation may be due to their extraordinary potency.  相似文献   

2.
Upon exposure of primary monolayer cultures of hepatocytes and H35 hepatoma cells, methptrexate (MTX) is taken up by carrier-mediated mechanisms and converted to γ-glutamyl derivatives with one to four residues being added. Under conditions that result in 90% or greater conversion, the primary metabolite in both cell types is MTX with three additional glutamates (4-NH2-10-CH3PteGlu4). When the time-dependent synthesis of MTX polyglutamates (4-NH2-10-CH3PteGlu2 and higher) at extracellular concentrations of 10 and 100 μm methotrexate is measured, both cell types exhibit linear synthesis for 4 to 6 hr, at which time an apparent steady state intracellular concentration of approximately 40 μm is reached. The concentration of MTX polyglutamate synthesized is not due a restriction in MTX since the hepatocytes and H35 cells accumulated 400 and 138 μm intracellular methotrexate, respectively, after 24 h in the presence of 100 μm extracellular MTX. Examination of MTX polyglutamate formation following a 24-h incubation showed concentration dependence with respect to intra- and extracellular MTX. Saturation was reached at a medium concentration of approximately 2 μm with both cell types which corresponded to 10 to 12 μm intracellular MTX. Placement of cells at steady state in medium lacking MTX results in the rapid equilibration of all free intracellular MTX with the medium. The MTX polyglutamates leave the cell by a slow loss of intact polyglutamates and also by intracellular cleavage to MTX followed by efflux. The longer-chain-length γ-glutamyl derivatives (Glu4–5) are more avidly retained by the cells than the shorter ones (Glu2–3).  相似文献   

3.
Cell chip was recently developed as a simple and highly sensitive tool for the toxicity assessment of various kinds of chemicals or nano-materials. Here, we report newly discovered potential cytotoxic effects of CdSe/ZnS quantum dots (QDs) on intracellular redox environment of neural cancer cells at very low concentrations which can be only detected by cell chip technology. Green (2.1 nm in diameter) and red (6.3 nm in diameter) QDs capped with cysteamine (CA) or thioglycolic acid (TA) were found to be toxic at 100 μg/mL when assessed by trypan blue and differential pulse voltammetry (DPV). However, in case of concentration-dependent cytotoxicity, toxic effects of TA-capped QDs on human neural cells were only measured by DPV method when conventional MTT assay did not show toxicity of TA-capped QDs at low concentrations (1-10 μg/mL). Red-TA QDs and Green-TA QDs were found to decrease electrochemical signals from cells at 10 μg/mL and 5 μg/mL, respectively, while cell viability decreased at 100 μg/mL and 50 μg/mL when assessed by MTT assay, respectively. The relative decreases of cell viability determined by MTT assay were 15% and 11.9% when cells were treated with 5-50 μg/mL of Red-TA QDs and 5-30 μg/mL of Green-TA QDs, respectively. However, DPV signals decreased 37.5% and 39.2% at the same concentration range, respectively. This means that redox environment of cells is more sensitive than other components and can be easily affected by CdSe/ZnS QDs even at low concentrations. Thus, our proposed neural cell chip can be applied to detect potential cytotoxicity of various kinds of molecular imaging agents simply and accurately.  相似文献   

4.
Studies were completed to characterize the cytotoxic and biochemical interaction of methotrexate (MTX) and 6-thioguanine (6-TG). Pretreatment of L1210 leukemia cells for 3 hr with MTX substantially enhanced the cytotoxicity of 6-TG. The LD90 of 6-TG in cells pretreated with 1μM MTX was 0.9pM, compared to an LD90 of 800 pM when the two drugs were given concurrently and an LD90 of 30 pM resulted with 6-TG alone. HPLC analysis of intracellular metabolites demonstrated an increased conversion of 6-TG to 6-TG-nucleotides in cells pretreated with MTX. A marked enhancement of 6-TG incorporation into RNA also was noted (MTX→6-TG: 350 fmol/μg RNA vs 6-TG: 98 fmol/μg RNA). However, there was a suppression of 6-TG incorporation into DNA when cells were pretreated with MTX (MTX→6-TG: 170 fmol/μg DNA vs 6-TG: 690 fmol/μg DNA). These results suggest that: 1) an enhancement of 6-TG antileukemic activity can be obtained with MTX pretreatment, and 2) the enhancement of 6-TG cytotoxicity following MTX exposure is not associated with 6-TG incorporation into DNA, but rather with incorporation of 6-TG into RNA. This drug sequence may be beneficial in the clinical treatment of leukemia.  相似文献   

5.
Harada H  Kamei Y 《Cytotechnology》1997,25(1-3):213-219
Extracts from 8 species of marine algae which showed selective cytotoxicity in our previous screening program, were further examined for cytotoxic spectra to five human leukemic cell lines. The extract from a red alga, Amphiroa zonata exhibited strong cytotoxicity to all human leukemic cell lines tested and murine leukemic cells L1210 at the final concentrations from 15 to 375 μg ml−1. Then the cytotoxicity was not found in normal human fibroblast HDF and murine normal cells NIH-3T3. The active extract fraction from this alga was soluble in higher polar organic solvents and water and heat-stable. The extract from a brown alga Dilophus okamurae with weak selective cytotoxic activity to L1210 cells exhibited not only strong cytotoxicity to L1210, but also to human leukemic cells, HL60 and MOLT-4 at 50 μg ml-1. While, the extract from a green alga, Cladophoropsis vaucheriaeformis with most selective cytotoxic activity, did not show cytotoxicity to any human leukemic cell lines tested at 50 μg ml-1. However, this extract showed strong cytotoxicity to two human leukemic cell lines and NIH-3T3 at 100 μg ml−1. Thus, it was considered that a red alga, Amphiroa zonata might be suitable natural source for development of anti-cancer agents without side-effect. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

6.
The aim of this study was to observe the cytotoxicity and apoptotic effects of nickel oxide nanoparticles on human cervix epithelioid carcinoma cell line (HeLa). Nickel oxide precursors were synthesized by an nickel sulphate-excess urea reaction in boiling aqueous solution. The synthesized NiO nanoparticles (<200 nm) were investigated by X-ray diffraction analysis and transmission electron microscopy techniques. For cytotoxicity experiments, HeLa cells were incubated in 50-500 μg/mL NiO for 2, 6, 12 and 16 hours. The viable cells were counted with a haemacytometer using light microscopy. The cytotoxicity was observed low in 50-200 μg/mL concentration for 16 h, but high in 400-500 μg/mL concentration for 2-6 h. HeLa cells' cytoplasm membrane was lysed and detached from the well surface in 400 μg/mL concentration NiO nanoparticles. Double staining and M30 immunostaining were performed to quantify the number of apoptotic cells in culture on the basis of apoptotic cell nuclei scores. The apoptotic effect was observed 20% for 16 h incubation.  相似文献   

7.
Granzyme B (GrB) is a key effector of cytotoxic lymphocyte-mediated cell death. It is delivered to target cells bound to the proteoglycan serglycin, but how it crosses the plasma membrane and accesses substrates in the cytoplasm is poorly understood. Here we identify two cationic sequences on GrB that facilitate its binding and uptake. Mutation of cationic sequence 1 (cs1) prevents accumulation of GrB in a distinctive intracellular compartment and reduces cytotoxicity 20-fold. Mutation of cs2 reduces accumulation in this intracellular compartment and cytotoxicity two- to threefold. We also show that GrB-mediated cytotoxicity is abrogated by heparin and that target cells deficient in cell surface sulfate or glycosaminoglycans resist GrB. However, heparin does not completely prevent GrB internalization and chondroitin 4-sulfate does not inhibit cytotoxicity, suggesting that glycosaminoglycans are not essential GrB receptors. We propose that GrB enters cells by nonselective adsorptive pinocytosis, exchanging from chondroitin sulfate on serglycin to anionic components of the cell surface. In this electrostatic "exchange-adsorption" model, cs1 and cs2 participate in binding of GrB to the cell surface, thereby promoting its uptake and eventual release into the cytoplasm.  相似文献   

8.
L1210/R81 lymphoma cells are resistant to methotrexate (MTX) by virtue of a 35-fold elevation in dihydrofolate reductase and an inability to transport the folate antagonist drug effectively. In a phosphate-containing buffer there was little or no influx into the resistant cells at either 1 or 50 μm MTX. Replacement of this buffer with a 4-(2-hydroxyethyl)-1-piperazine-N′-2-ethanesulfonic acid-Mg2+ system resulted in an apparent influx of MTX into the resistant cells. Under these conditions, L1210/R81 cells achieved an apparent steady state at an extracellular MTX concentration of 50 μm. The apparent steady-state level of 5 nmol [3H]MTX109 cells was well below the intracellular level of dihydrofolate reductase (45 nmol/109 cells). Efflux experiments at the apparent steady state indicated that 60% of the MTX was very rapidly removed from the cells by washing. Over the range of the experiment a further 20% of the MTX effluxed more slowly (t12 = 12 min). The apparent influx into the resistant cells at 5 μm MTX was inhibited 13% by sodium azide (100 μm) and initially stimulated, then inhibited, by p-chloromercuriphenylsulfonic acid (100 μm). 5-Methyltetrahydrofolate (100 μm) had little effect on the process while aminopterin (100 μm) was inhibitory (68%). Kt and V values of 2 × 10?5m and 0.31 nmol [3H]MTX109 cells/min, respectively, were determined for the apparent influx in L1210R81 cells. Comparison of apparent MTX influx in the resistant cells with MTX transport in the sensitive cells indicates profound differences in the two processes. The evidence suggests that the apparent influx in the former cell line may consist of MTX binding to the cell membrane together with a small degree of MTX influx into the intracellular compartment.  相似文献   

9.
Summary In vitro effects of methotrexate (MTX) on interleukin-2(IL-2)-mediated cytotoxicity of peripheral blood mononuclear cells (PBMC) were studied. PBMC were incubated with human recombinant IL-2 (25 U/ml) for 72 h; during the last 24 h, various concentrations (10 pM–1 µM) of MTX were added to the culture. Cytotoxicity against k562 cells was measured by a 4-h51Cr-release assay. The IL-2-mediated cytotoxicity was paradoxically increased at around a concentration (10 nM) MTX. Such a low concentration of MTX showed no anti-proliferative effect on cell growth. This enhancement with 10 nM MTX was shown only in an E-rosette+ (E+) population, but not in E-rosette (E). In addition, when E+ cells were treated with an anti-CD16 monoclonal antibody plus complement after incubation with IL-2 and MTX, MTX-induced enhancement was lost, suggesting that an E+CD16+ cell population was mainly involved in this augmentation. Positively sorted E+CD16+ cells showed similar enhancement of cytotoxicity after treatment with IL-2 plus MTX. On the other hand, MTX treatment did not show the phenotypical changes including of the E+CD16+ cells, indicating that this treatment did not affect the differentiation and proliferation of the specific cell subset. Our results indicate that a low dose of MTX could have a role in the regulation of immunological anti-cancer surveillance systems through the natural killer and lymphokine-activated cytotoxic cells.This work was supported in part by a Grant-in-Aid for Cancer Research (1–10) from the Ministry of Health and Welfare in Japan  相似文献   

10.
Previous studies have suggested that heparin-like glycosaminoglycans may be endogenous inhibitors of smooth muscle proliferation in the vessel wall. The purpose of this study was to determine the effects of exogenous glycosaminoglycans on rat vascular (aortic) smooth muscle cell migration following wounding in vitro. Our data indicate that heparin and related molecules (iota carrageenan, dextran sulfate), but not other glycosaminoglycans (hyaluronate, chondroitin, and dermatan sulfates), inhibit smooth muscle cell motility in a cell-specific, dose-dependent, and reversible fashion. The effect of heparin was maximal (60% inhibition) at 10 μg/ml; a half-maximal effect was observed at 1 μg/ml; Heparin did not significantly affect the migration of bovine aortic endothelium or Swiss 3T3 cells. These observations support the concept that heparin-like glycosaminoglycans may be important regulators of vascular smooth muscle cell function.  相似文献   

11.
Fifteen diterpenoids ( 1 – 15 ), including three undescribed ones with ent‐atisane skeleton, eupnerias G–I ( 1 – 3 ), were obtained from Euphorbia neriifolia. Compounds 1 – 3 were established through comprehensive spectroscopic analysis. Compounds 4 and 5 exhibited obvious anti‐HIV‐1 effect, and their EC50 were 6.6±3.2 and 6.4±2.5 μg mL?1, respectively. Compound 6 exhibited moderate cytotoxicity on HepG2 and HepG2/Adr cells with IC50 at 13.70 and 15.57 μm , respectively. In addition, compound 15 exhibited significant cytotoxicity on HepG2 cell lines (IC50=0.01 μm ), while it did not show any cytotoxicity against HepG2/Adr cell lines.  相似文献   

12.
Based on the concern of organochlorides in the environment and in human tissue, this study was designed to determine whether various noncytotoxic levels of heptachlor and heptachlor epoxide could inhibit, reversibly, gap junctional intercellular communication in human breast epithelial cells (HBEC). Cytotoxicity and gap junctional intercellular communication (GJIC) were evaluated by lactate dehydrogenase assay and fluorescence redistribution after photobleaching analysis, respectively. Both heptachlor and heptachlor epoxide were noncytotoxic up to 10 μg/ml. At this concentration, heptachlor and heptachlor epoxide inhibited GJIC of normal human breast epithelial cells after 1 h treatment. Within a 24 h treatment with heptachlor and heptachlor epoxide at 10 μg/ml, recovery of GJIC had not returned. GJIC completely recovered after a 12 h treatment of 1 μg/ml heptachlor epoxide, but it did not recover after a 24 h treatment of 1 μg/ml heptachlor. RT-PCR and Western blots were analyzed to determine whether the heptachlor or heptachlor epoxide might have altered the steady-state levels of gap junction mRNA and/or connexin protein levels or phosphorylation state. No significant difference in the level of connexin 43 (Cx43) message between control and heptachlor-treated cells was observed. Western blot analyses showed hypophosphorylation patterns in cells treated with 10 μg/ml heptachlor and heptachlor epoxide for 1 h with no recovery within 24 h. Immunostaining of Cx43 protein in normal HBEC indicated that heptachlor and heptachlor epoxide caused a loss of Cx43 from the cell membranes at noncytotoxic dose levels. Taken together, these results suggest that heptachlor and heptachlor epoxide can alter GJIC at the post-translational level, and that, under the conditions of exceeding a threshold concentration in the breast tissue containing ‘initiated’ cells for a long time and not being counteracted by anti-tumor-promoting chemicals, they could act as breast tumor promoters.  相似文献   

13.
Adenosine (10 μM) stimulates the initial growth rate of BHK/21 cells seeded at low but not high density in monolayer culture; it does not affect final cell density or permit growth in agar. In labelling experiments With tritiated thymidine, adenosine also increases the response of quiescent cells to low concentrations of serum. Dialysis of serum to remove oxypurines only marginally reduces its effect on quiescent cell labelling or growth, indicating that BHK/21 cells are able to synthesise purines. The response of quiescent cells to 5% serum is inhibited by high MW (2 × 106) dextran sulphate at 2 μg per milliliter. Low MW dextran sulphate (30,000) and heparin at 20 μg per milliliter produce the same effect. Exogenous adenosine (10 μM) prevents this inhibition. Many other purine derivatives replace adenosine for all the above activities but xanthine is completely inactive in all. It, therefore, appears that nucleotide synthesis is a necessary function of these compounds. The growth of cells of a polyoma-virus-transformed BHK/21 line in monolayer is not stimulated by exogenous purine, though their colony-forming ability in agar is increased five-fold. The stimulating effects of exogenous purines on normal BHK/21 cells and the absolute requirement for them in the presence of polyanions is discussed in relation to possible mechanisms of growth control.  相似文献   

14.
Conjugation of boron nanoparticles with porphyrins is an attractive way to create dual agents for anticancer boron neutron capture therapy (BNCT) and photodynamic therapy (PDT). Properties of chlorin e(6) conjugated with two cobalt bis(dicarbollide) nanoparticles (1) or with a closo-dodecaborate nanoparticle (2) are reported. Fluorescent dianionic conjugates 1 and 2 penetrate in A549 human lung adenocarcinoma cells, stain cytoplasm diffusely and accumulate highly in lysosomes but are not toxic themselves for cells. Average cytoplasmic concentration of boron atoms (B) achieves 270 μM (ca. 2 × 10(8) B/cell) and 27 μM (ca. 2 × 10(7) B/cell) at the 1.5 μM extracellular concentration of 1 and 2, respectively, that makes conjugate 1 especially suitable for BNCT. Conjugate 2 causes photoinduced cell death at micromolar concentrations and can be considered also as a photosensitizer for PDT. Conjugates 1 and 2 have high quantum yields of singlet oxygen generation (0.55 and 0.85 in solution, respectively), identical intracellular localization and similar lipid-like microenvironment but conjugate 1 possesses no photoinduced cytotoxicity. A presence of cobalt complexes in conjugate 1 is supposed to be a reason of the observed antioxidative effect in cellular environment, but an exact mechanism of this intriguing phenomenon is unclear. Due to increased intracellular accumulation and absence of photoinduced cytotoxicity conjugate 1 is promising for fluorescence diagnostics of cancer.  相似文献   

15.
The cytochalasin B (CYB)-blocked binucleated cell assay has been explored to analyze munuclei and cell cycle kinetics using 2 known mutagenic carcinogens in V79 Chinese hamster lung cells. To determine the optimum time to obtain the maximum number of binucleated cells for munucleus analysis, duplicate cultures of exponentially growing cells were treated with 3 μg/ml CYB for varying durations (8–48 h). A peak appearance of binucleated cells at 16 h in the presence of CYB suggested this as an optimum time for munucleus analysis in binucleated V79 cells. To evaluate the capacity for induction of munuclei in V79 cells, 2 mutagenic carcinogens, mitomycin C (0.125–1.0 μg/ml) and cyclophosphamide (2–12 μg/ml) were tested in duplicate cultures. Mitomycin C, a direct-acting alkylating agent, caused approximately an 18-fold increase in munucleus frequency over controls at the highest concentration tested (1.0 μg/ml), and this increase occurred in a dose-related manner (r = 0.92). The concentrations of mitomycin C tested also caused a significant dose-related cell cycle delay, thus suggesting cytotoxicity to V79 cells. Cyclophosphamide, an indirect-acting alkylating agent, requiring the presence of S9 mix, caused approximately a 17-fold increase in munucleus frequency over controls at the highest tested concentration (12 μg/ml), with a clear dose response (r = 0.99). The various concentrations of cyclophosphamide also caused cytotoxicity in a dose-related fashion. Thus, this study demonstrates the usefulness of the cytokinesis-block method in V79 cells as a possible screen to analyze munucleus induction and cytotoxicity. Because this approach is much less labor intensive than conducting a structural chromosomal analysis, this assay has great potential both as an initial screen for clastogenic activity and as a tool for investigating the underlying mechanisms for clastogenicity.  相似文献   

16.
To compare the cytotoxicity on HeLa cells induced by nanosized and microsized tellurium powders, HeLa cells were exposed to different concentrations of tellurium powders (0, 50, 100, 150 and 200 μg/mL) for 12 h. In this study, detection of a series of biomarkers, including reactive oxygen species (ROS), glutathione (GSH), 8-hydroxy-2′-deoxyguanosine (8-OHdG), in addition to DNA and protein crosslink (DPC) and MTTassay, were conducted to evaluate the cytotoxicity. It is indicated that compared with the control group, there was no significant difference in the induced cytotoxicity at concentrations lower than 50 μg/mL for both nanosized and microsized tellurium powders. While there appears a significant difference in the induced cytotoxicity for nanosized tellurium powders when the concentration is higher than 100 μg/mL as well as for microsized tellurium powders when the concentration is higher than 200 μg/mL. Moreover, it is found that the cytotoxicity induced on HeLa cells exhibits a certain dose-effect relationship with the concentration of tellurium powders. A conclusion has been reached that the toxicity on HeLa cells can be induced by both nanosized and microsized tellurium powders, and the toxicity of the nanosized tellurium powders is significantly greater than the microsized one.  相似文献   

17.
Heparin specifically and saturably binds to bovine spermatozoa and stimulates capacitation as assessed by the ability of spermatozoa to undergo a zona pellucida-induced acrosome reaction (AR) in vitro. However, the structural features of heparin important for capacitation are poorly understood The purpose of this study was to determine the importance of the sulfatc content of heparin for its potency to bind to bull spermatozoa and promote agglutination and capacitation. The pyridine salt of heparin was Nndesulfated, which reduced its mean sulfate content from 19.7% to 11.6%. The N-desulfated heparin was then resulfated by incubation with trimcthylamine sulfur trioxide for 6,12, or 24 hr, raising sulfate to original concentrations. Heparin but not N-desulfated beparin competed with [3H]-heparin to bind to spermatozoa. Heparin at 11.6 μg/ml reduced [3H]-heparin binding by half when competing with a saturating concentration of the radidabeled compound (12 μg/ml). N-desulfated heparin did not displace [3H]-heparin. Heparin, resulfated 6 hr or 12 hr, was equal to native heparin in binding potency. Heparin at 50,100, or 250 μg/ml caused more than 40% of the cells to head-to-head agglutinate in aggregates of 8 or more. N-desulfated heparin did not cause agglutination. After spermatozoa were incubated with 0, 5,10, 50, 100, or 250 μg/ml of heparin for 4 hr, 100 μg/ml of lysophosphatidylcholine (LPC)-induccd AR within 20 min in 21.3, 37.7, 27.8, 45.3, 54.2, or 42.5% of the cells, respectively. Sperm exposed to the same concentrations of N-dcsulfated heparin exhibited AR of 17.7,27.3,24.3,22.5,27.7, or 33.8%, respectively, following exposure to LPG Resulfated heparin did not agglutinate or capacitate spermatozoa. In conclusion, N-desulfation of heparin abolished heparin's ability to bind to, agglutinate, and capacitate bovine spermatozoa. Resulfation of N-desulfated heparin restored binding activity but not agglutination or capacitation activity.  相似文献   

18.
Heparin inhibited the adenylate cyclase activity of semipurified rat pancreatic plasma membranes stimulated by hormones and by Gpp(NH)p but not by fluoride or when in the persistently active state. When observed, the inhibition was rapid and sustained. It was of a noncompetitive type and never exceeded 20% for secretin. The inhibition of Gpp(NH)p-stimulated activity was more pronounced (48% inhibition at a heparin concentration of 50 μg/ml). For the C-terminal octapeptide of pancreozymin (CCK-8)-stimulated adenylate cyclase, the inhibition amounted to 93% at 50 μg/ml. This inhibition was competitive at low heparin concentration and of a mixed type above 10 μg/ml. Besides, heparin inhibited (I50 = 6 μg/ml) the binding of peptides of the CCK family to their specific receptors without affecting the apparent Kd value of binding. Taken together, these relatively specific effects of heparin gave evidence in favor of the existence of CCK spare receptors. Dextran sulfate was more potent than heparin as an inhibitor of adenylate cyclase activation while chondroitin-4-sulfate and chondroitin-6-sulfate were ineffective. Dansylated pancreatic plasma membranes exhibited characteristics of adenylate cyclase activation by CCK-8 which were similar to those found for untreated membranes exposed to heparin.  相似文献   

19.
The effect of heparin on the in vitro lysis of EL4 tumor cells by immune BALB/C lymphocytes was investigated by using a 51Cr-release cytotoxicity. assay. Powdered heparin did not inhibit lymphocyte-mediated cytotoxicity (LMC) at concentrations up to 500 units/ml. Heparin containing 9 mg/ml of benzyl alcohol (BA), as preservative, significantly reduced the LMC. BA alone at 1 mg/ml inhibited LMC without any apparent toxic effect on the target cells or on the immune lymphocytes. The inhibitory effect of three different heparin preparations was found to be related to the BA concentration rather than the amount of heparin. However, low concentrations of heparin (0.5 or 1 unit/ml) significantly enhanced the LMC. Our findings are in contrast to previous reports suggesting a depressive effect of heparin on LMC.  相似文献   

20.
While Cry1Ac has been known to bind with larval midgut proteins cadherin, APN (amino peptidase N), ALP (alkaline phosphatase) and ABCC2 (adenosine triphosphate‐binding cassette transporter subfamily C2), little is known about the receptors of Cry2Ab. To provide a clue to the receptors of Cry2Ab, we tested the baseline cytotoxicity of activated Cry1Ac and Cry2Ab against the midgut and fat body cell lines of Helicoverpa zea and the ovary cell line of Spodoptera frugiperda (SF9). As expected, the descending order of cytotoxicity of Cry1Ac against the three cell lines in terms of 50% lethal concetration (LC50) was midgut (31.0 μg/mL) > fat body (59.0 μg/mL) and SF9 cell (99.6 μg/mL). By contrast, the fat body cell line (LC50 = 7.55 μg/mL) was about twice more susceptible to Cry2Ab than the midgut cell line (16.0 μg/mL), the susceptibility of which was not significantly greater than that of SF9 cells (27.0 μg/mL). Further, ligand blot showed the binding differences between Cry1Ac and Cry2Ab in the three cell lines. These results indicated that the receptors of Cry2Ab were enriched in fat body cells and thus largely different from the receptors of Cry1Ac, which were enriched in midgut cells.  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