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.
Stable and heritable variants of Chinese hamster ovary (CHO) cells which are resistant to different levels (0.1, 1.0 and 10 μg/ml) of the toxin abrin have been isolated and characterized. The frequency of resistant colonies to abrin was increased with the concentration of a chemical mutagen. There was no effect of cell density or cross-feeding on the recovery of variants. In experiments using fluorescein-labeled abrin and ricin which bind to terminal (non-sialylated) galactose residues of cell-surface oligosaccharides, parental cells exhibited strong binding toward both toxins, whereas no fluorescence was observed in the resistant clones. A fluorescein-conjugated lectin, BS II, which is specific for terminal N-acetyl- -glucosaminyl residues, did not interact with the parental cells, but did with the resistant clones. This suggests that on the surface of resistant cells the number of terminal galactosyl residues of oligosaccharide chains in glycoproteins was reduced, exposing the penultimate N-acetyl- -glucosaminyl residues. The number of available endogenous acceptor sites for galactosyl transferase in the abrin-resistant clones was directly proportional to the degree of resistance. In the presence of great excess of exogenous acceptor, the rates of galactosyl transfer were similar in all the abrin-resistant cell types tested, with levels ranging from 1.4 to 1.7 times parental cell values. Studies with tetraploid cell hybrids reveal that resistance was a recessive trait. Fluctuation analysis showed that abrin resistance occurred in CHO cell populations at a rate of 4−7 × 10−8/cell/generation. The system may serve as a new marker for quantitative mutagenesis studies.  相似文献   

2.
Six embryonal carcinoma (EC) cell lines that are resistant to the cytotoxic, galactose-specific lectin abrin were isolated from mutagenized populations of either PSA-1 or F9 cells. The LD10 for each of the variant lines was at least 150-fold greater than that for parental cells. Indirect cytotoxicity tests demonstrated that all of the variant cell lines lacked both Stage Specific Embryonic Antigen-1 (SSEA-1, less than 1% of wild-type levels) and Forsmann antigen (less than 5% of wild-type levels). When abrin-resistant cells were fused to previously isolated SSEA-1-negative cells (M. J. Rosenstraus (1983), Dev. Biol. 99, 318-323) that express Forsmann antigen, the resulting hybrids expressed SSEA-1. This implies the mutation conferring abrin resistance is in a different gene than that defined by the previously isolated mutation. Thus, we have identified two genes that are required for SSEA-1 expression, one of which also appears to be required for Forsmann antigen expression. The F9-derived variants differentiated into visceral-like or parietal-like endoderm when treated with retinoic acid in the absence or presence of 8-bromo-cAMP, respectively. PSA-1-derived variants formed differentiated teratocarcinomas containing derivatives of all three germ layers. Thus the SSEA-1 and Forsmann haptenic determinants are not required for EC cells to differentiate into a broad spectrum of cell types; nor do they appear to be involved in the cell-cell interactions that are postulated to regulate visceral versus parietal endoderm differentiation.  相似文献   

3.
Previous studies with the multidrug-resistant human HL60 cell line have shown a 3-4-fold decrease in VP-16 accumulation compared to the sensitive cell line, while the degree of resistance to VP-16 was 300-fold, indicating that other mechanisms of resistance are also operative. Since VP-16 has been shown to interfere with topoisomerase II activity, we have evaluated VP-16-dependent DNA strand break formation in the drug-sensitive and -resistant HL60 cells. Studies reported here show that the drug-resistant HL60 cells are extremely resistant to VP-16-dependent DNA cleavage compared to the sensitive cells. This decrease in DNA cleavage activity in the presence of VP-16 was, in part, related to a 2-3-fold decrease in both the amount and activity of topoisomerase II in the resistant cell line compared to the sensitive cells. Nuclei from the resistant cell line were markedly more resistant to VP-16-dependent DNA cleavage than the WT cell nuclei. Interestingly, WT nuclei were found to be relatively more resistant to VP-16-induced DNA cleavage than the intact WT cells. Addition of WT cytosolic proteins to WT nuclei, however, significantly stimulated VP-16-dependent DNA cleavage and slightly increased DNA cleavage in resistant cell nuclei. In contrast, cytosolic proteins from the resistant cells had no effect on DNA cleavage in nuclei isolated from either cell line. These observations indicate that a decrease in the amount and activity of topoisomerase II in resistant HL60 cells translates into a decrease in VP-16-dependent DNA breakage and contributes to the resistance to VP-16. Furthermore, the cytosolic fraction from WT cells contains some factor, not present in the resistant cells, which is necessary for the maximal drug-induced DNA cleavage.  相似文献   

4.
Ricin, a toxic lectin from Ricinus communis, is composed of two different polypeptide chains, A and B, and the ricin A chain (RA) blocks protein synthesis. We studied cell lines resistant to cytotoxic action of RA. One low-RA-resistant cell line, AR10, isolated from Chinese hamster ovary (CHO) cells, was resistant to a low dose of RA (1 microgram/ml) and showed a 10-fold-higher resistance to RA and ricin than that of CHO. We further mutagenized AR10 to isolate high-RA-resistant cell lines AR100-6, AR100-9, and AR100-13, which were resistant to higher doses of RA and ricin (100- to 1,000-fold) than CHO was. The binding of [125I]ricin to AR10, AR100-6, AR100-9, and AR100-13 cells was decreased to about 30% of that of CHO. The internalization of [125I]ricin in AR10 cells and in the high-RA-resistant clones was the same. Polyuridylate-dependent polyphenylalanine synthesis, using S-30 extracts from either AR100-9 or AR100-13, was about 100-fold more resistant to the inhibitory action of RA than when CHO, AR10, and AR100-6 cells extracts were used. The protein synthesis with ribosomes (80S) from AR100-9 or AR100-13 was 10- to 100-fold more resistant to RA than it was with parental ribosomes when combined with the S-100 fraction of CHO cells. The polyphenylalanine synthesis assay using the ribosomes constituted from the 60S subunit of AR100-9 and the 40S subunit of CHO indicated that the resistant phenotype of AR100-9 cells is due to an alteration of the 60S ribosomal subunit.  相似文献   

5.
6.
Abrin B chain and trypsin inhibitor isolated from Acacia confusa (ACTI) were covalently linked to form a chimeric protein (ANB-ACTI) with N-succinimidyl-3-(-2-pyridyldithio)propionate. The chimeric protein had 31% of trypsin inhibitory activity of ACTI and 7% of hemagglutinating activity of abrin B chain, but no inhibition on protein biosynthesis. ANB-ACTI had strong inhibitory effects on the growth of sarcoma 180 cells and Hela cell culture while the mixture of an equivalent amount of free abrin B chain and ACTI did not. The results suggests that abrin B chain of chimeric protein may act as a vector to carry ACTI into the tumor cells. ACTI into the tumor cells. ACTI in the chimeric protein potentiates its antitumor activity as well as its resistance to tryptic digestion.  相似文献   

7.
Previous studies with the multidrug-resistant human HL60 cell line have shown a 3–4-fold decrease in VP-16 accumulation compared to the sensitive cell line, while the degree of resistance to VP-16 was 300-fold, indicating that other mechanisms of resistance are also operative. Since VP-16 has been shown to interfere with topoisomerase II activity, we have evaluated VP-16-dependent DNA strand break formation in the drug-sensitive and -resistant HL60 cells. Studies reported here show that the drug-resistant HL60 cells are extremely resistant to VP-16-dependent DNA cleavage compared to the sensitive cells. This decrease in DNA cleavage in the of VP-16 was, in part, related to a 2–3-fold decrease in both the amount and activity of topisomerase II in the resistant cell line compared to the sensitive cells. Nuclei from the resistant cell line were markedly more resistant to VP-16-dependent DNA cleavage than the WT cell nuclei. Interestingly, WT nuclei were found to be relatively more resistant to VP-16-induced DNA cleavage than the intact WT cells. Addition of WT cytosolic proteins to WT nuclei, however, significantly stimulated VP-16-dependent DNA cleavage and slightly increased DNA cleavage in resistant cell nuclei. In contrast, cytosolic proteins from the resistant cells had no effect on DNA cleavage in nuclei isolated from either cell line. These observations indicate that a decrease in the amount and activity of topoisomerase II in resistant HL60 cells translates into a decrease in VP-16-dependent DNA breakage and contributes to the resistance to VP-16. Furthermore, the cytosolic fraction from WT cells contains some factor, not present in the resistant cells, which is necessary for the maximal drug-induced DNA cleavage.  相似文献   

8.
Saxena PK  King J 《Plant physiology》1988,86(3):863-867
Cells resistant to the sulfonylurea herbicides chlorsulfuron and sulfometuron methyl were isolated from a predominantly haploid cell suspension culture of Datura innoxia P. Mill. Exponentially growing cell colonies (aggregates of about 40 cells) were mutagenized with ethyl methane sulfonate, subcultured for 10 days to allow growth recovery and plated on a medium containing either chlorsulfuron or sulfometuron methyl at a concentration (10−8 molar) which killed wild type cells. Surviving clones were picked up after 3 to 4 weeks, further proliferated as callus or cell suspension cultures, and tested for their resistance to both the sulfonylureas and imidazolinones, a chemically different class of herbicides. The variants were stable and showed high (100- to 1000-fold) resistance to the sulfonylureas. While some also exhibited cross resistance to imidazolinones, others showed no cross-resistance at all or, as in one case, greater sensitivity than wild type cells to the imidazolinones. Both classes of herbicides tested inhibited acetolactate synthase activity isolated from wild type cells. The acetolactate synthase of the resistant variants, however, was found to be resistant to the sulfonylureas and also to the imidazolinone(s) in those cells showing cross-resistance to the latter. The lack of cross-resistance observed in some cases provides evidence that the two groups of herbicides have slightly different sites on the acetolactate synthase molecule.  相似文献   

9.
The toxic lectin modeccin, which inhibits protein synthesis in eukaryotic cells, is cleaved upon treatment with 2-mercaptoethanol into two peptide chains which move in polyacrylamide gels at rates corresponding to molecular weights 28,000 and 38,000. After reduction, the toxin loses its effect on cells, while its ability to inhibit cell-free protein synthesis increases. Like abrin and ricin it inhibits protein synthesis by inactivating the 60S ribosomal subunits. Modeccin binds to surface receptors containing terminal galactose residues. Competition experiments with various glycoproteins indicate that the modeccin receptors are different from the abrin receptors. In addition, they were present on HeLa cells in much smaller numbers. Moreover, mutant lines resistant to abrin and ricin were not resistant to modeccin and vice-versa. The toxin resistance of various mutant cell lines could not be accounted for by a reduced number of binding sites on cells. The data are consistent with the view that the cells possesss different populations of binding sites with differences in ability to facilitate the uptake of the toxins and that in the resistant lines the most active receptors have been reduced or eliminated.  相似文献   

10.
B P Kopnin  A A Stavrovskaia 《Genetika》1979,15(12):2233-2236
The studies of the high level of colchicine resistance of mouse L cells have shown that two mutagens (EMS and NMM) do not induce cell variants resistant to 8 microgram/ml of colchicine in the population of mouse heteroploid L-53 cells (subline of L cells, the level of colchicine resistance 140) and that colchicine resistance of L-53 cells gradually diminishes when cells are propagated in non-selective conditions: after 1 month it diminishes 2-fold, after 3 month--9-fold. The extent of the decrease of the drug resistance was the same in 6 independent cultures obtained from the inoculum of 200 cells and in control cultures propagated by large quantities of cells. These data coincide with the results of the previous studies of lower level of colchicine resistance. In both studies the frequency of the occurrence of colchicine resistant variants in selective medium was about 2.10(-4). These data are consistent with the hypothesis that colchicine resistance of mouse L cells is not due to a gene mutation.  相似文献   

11.
Therapeutic monoclonal antibodies (mAb) are often produced in Chinese hamster ovary (CHO) cells. Three commonly used CHO host cells for generating stable cell lines to produce therapeutic proteins are dihydrofolate reductase (DHFR) positive CHOK1, DHFR‐deficient DG44, and DUXB11‐based DHFR deficient CHO. Current Genentech commercial full‐length antibody products have all been produced in the DUXB11‐derived DHFR‐deficient CHO host. However, it has been challenging to develop stable cell lines producing an appreciable amount of antibody proteins in the DUXB11‐derived DHFR‐deficient CHO host for some antibody molecules and the CHOK1 host has been explored as an alternative approach. In this work, stable cell lines were developed for three antibody molecules in both DUXB11‐based and CHOK1 hosts. Results have shown that the best CHOK1 clones produce about 1 g/l for an antibody mAb1 and about 4 g/l for an antibody mAb2 in 14‐day fed batch cultures in shake flasks. In contrast, the DUXB11‐based host produced ~0.1 g/l for both antibodies in the same 14‐day fed batch shake flask production experiments. For an antibody mAb3, both CHOK1 and DUXB11 host cells can generate stable cell lines with the best clone in each host producing ~2.5 g/l. Additionally, studies have shown that the CHOK1 host cell has a larger endoplasmic reticulum and higher mitochondrial mass. © 2013 American Institute of Chemical Engineers Biotechnol. Prog., 29:980–985, 2013  相似文献   

12.
The calcium channel blockers verapamil and diltiazem have been shown to reverse multidrug resistance, but the mechanism of action of these agents is still unknown. We measured [3H]verapamil, [3H]desmethoxyverapamil, [3H]diltiazem, and [3H]nitrendipine binding to membrane vesicles made from drug-sensitive (KB-3-1), multidrug-resistant (KB-C4 and KB-V1), and revertant (KB-V1-R2) cells. Membrane vesicles from KB-V1 cells bound 10-20-fold more [3H]verapamil and [3H]diltiazem and about 30-fold more [3H]desmethoxyverapamil than did vesicles from the parental KB-3-1 or revertant KB-V1-R2 cell lines. These drugs reverse the multidrug resistance phenotype by increasing accumulation of drugs in the resistant cells. No difference in binding of [3H]nitrendipine, which did not reverse drug resistance, was observed. The binding of vinblastine, desmethoxyverapamil, and diltiazem to KB-V1 vesicles was specific and saturable and was inhibited by desmethoxyverapamil and quinidine greater than vinblastine and diltiazem much greater than daunomycin. In addition, verapamil and diltiazem inhibited the vinblastine photoaffinity labeling of P170, the protein previously shown to be a marker of multidrug resistance.  相似文献   

13.
Following transfection of genomic mouse DNA into an SV40 transformed fibroblast cell line from a patient with Xeroderma pigmentosum (complementation group A, XPA), a single UV resistant cell clone was isolated out of a total of 10(4) independent transfectants. The recipient XPA cell line has as yet not produced spontaneous revertants among 2.2 x 10(8) cells. The isolated cell clone contains 50-70 kb of mouse sequences which are heavily amplified (500-fold), and has acquired both intermediate resistance to UV killing and intermediate unscheduled DNA synthesis (UDS) capacity. By continued passage without selective pressure, cells were generated, which had lost both the dominant marker gene and repetitive mouse sequences. Single colonies of these cells were still intermediately resistant to UV suggesting that either undetected unique mouse DNA had segregated from the bulk of repetitive DNA, or, more likely, that the initially isolated transfectant was a spontaneous revertant. This documents that a persuasive clone isolated can still be a false positive (spontaneous revertant) and that an extremely laborious approach may lead into a dead end.  相似文献   

14.
A line of human lymphocytic leukemia cells (CCRF-CEM) has been obtained which is 140-fold resistant to the potent cell growth inhibitor 5-fluoro-2'-deoxyuridine (FdUrd). The cells were also 11-fold cross-resistant to 5-fluorouracil. In contrast to several previous studies involving FdUrd-resistant mouse cells, thymidylate synthetase levels were not substantially elevated in these FdUrd-resistant human leukemic cells. Thymidine kinase activity was also unchanged in the resistant cells, although the levels of 5-fluoro-2'-deoxyuridylate (FdUMP), the potent inhibitor of thymidylate synthetase, generated at equimolar doses of FdUrd were about 40% lower than in the sensitive cells. Studies of the kinetics of FdUMP binding to thymidylate synthetase isolated from the FdUrd-resistant cells disclosed a considerably higher dissociation constant (Kd = 1.0 X 10(-9) M) for the ternary covalent enzyme . FdUMP . 5,10-methylene tetrahydrofolate complex compared to the value obtained with enzyme from sensitive cells (Kd = 4.4 X 10(-11) M). The thymidylate synthetase from the FdUrd-resistant cells also showed 17-fold weaker binding of 2'-deoxyuridylate, even though the Km value for 2'-deoxyuridylate was 3-fold lower compared to the enzyme from FdUrd-sensitive cells. The turnover number of the altered enzyme was 1.8-fold higher than that for the normal enzyme but the rate constants for the release of FdUMP from the ternary complex, which is also an enzyme-catalyzed reaction, were identical for both enzymes. Electrophoresis of the radiolabeled ternary complexes on nondenaturing gels showed small but reproducible differences in migration rates. These results demonstrate that the mechanism of resistance to FdUrd in this cell line involves an alteration in the target enzyme, thymidylate synthetase, which causes it have a lower affinity for nucleotides.  相似文献   

15.
There is a desire to engineer mammalian host cell lines to improve cell growth/biomass accumulation and recombinant biopharmaceutical protein production in industrially relevant cell lines such as the CHOK1 and HEK293 cell lines. The over-expression of individual subunits of the eukaryotic translation factor eIF3 in mammalian cells has previously been shown to result in oncogenic properties being imparted on cells, including increased cell proliferation and growth and enhanced global protein synthesis rates. Here we report on the engineering of CHOK1 and HEK cells to over-express the eIF3i and eIF3c subunits of the eIF3 complex and the resultant impact on cell growth and a reporter of exogenous recombinant protein production. Transient over-expression of eIF3i in HEK293 and CHOK1 cells resulted in a modest increase in total eIF3i amounts (maximum 40% increase above control) and an approximate 10% increase in global protein synthesis rates in CHOK1 cells. Stable over-expression of eIF3i in CHOK1 cells was not achievable, most likely due to the already high levels of eIF3i in CHO cells compared to HEK293 cells, but was achieved in HEK293 cells. HEK293 cells engineered to over-express eIF3i had faster growth that was associated with increased c-Myc expression, achieved higher cell biomass and gave enhanced yields of a reporter of recombinant protein production. Whilst CHOK1 cells could not be engineered to over-express eIF3i directly, they could be engineered to over-express eIF3c, which resulted in a subsequent increase in eIF3i amounts and c-Myc expression. The CHOK1 eIF3c engineered cells grew to higher cell numbers and had enhanced cap- and IRES-dependent recombinant protein synthesis. Collectively these data show that engineering of subunits of the eIF3 complex can enhance cell growth and recombinant protein synthesis in mammalian cells in a cell specific manner that has implications for the engineering or selection of fast growing or high producing cells for production of recombinant proteins.  相似文献   

16.
The mechanism for cisplatin resistance in cisplatin-resistant KCP-4 cells was studied. Although multidrug resistance-associated protein (MRP) was not detected in KCP-4 cells, the cells were more resistant to heavy metals than multidrug-resistant C-A120 cells that overexpressed MRP. KCP-4 cells expressed metallothionein, but it was scarcely involved in cisplatin resistance in these cells. KCP-4 cells did not express canalicular multispecific organic anion transporter (cMOAT). The glutathione(GSH) level was 4.7-fold higher in KCP-4 cells than in KB-3-1 cells. When the GSH level in KCP-4 cells was decreased by treating the cells with buthionine sulfoximine and nitrofurantoin, the accumulation of and sensitivity to cispaltin in the cells were increased. C-A120 cells were only 3.0-fold more resistant to cisplatin than KB-3-1 cells and this resistance was not affected by the increased glutathione level. The accumulation of platinum in C-A120 and KCP-4 cells was 68.5 and 20.4% of that in KB-3-1 cells, respectively, while the intracellular levels of antimony potassium tartrate in C-A120 and KCP-4 cells were 13.2 and 9.9% of that in KB-3-1 cells, respectively. The ATP-dependent efflux of antimony was enhanced in both C-A120 and KCP-4 cells. These results, taken together, suggest an efflux pump for heavy metals different from MRP and cMOAT is involved in cisplatin resistance in KCP-4 cells.  相似文献   

17.
The effects of the plant toxin abrin on normal mouse embryonic fibroblasts (MEF), an untransformed mouse cell line (NIH 3T3), and two mouse tumor cell lines (LMTK- and S-180) were studied. Measurements of cell growth and colony formation showed that MEF and S-180 cells were more sensitive to abrin intoxication than NIH 3T3 and LMTK- cells. Also, the effects of abrin on the inhibition of [3H]leucine and [3H]thymidine incorporation were more evident in MEF and S-180 cells. The basis for these varying responses to abrin by the four different cells was examined. The number of abrin binding sites per cell was determined from [125I]abrin binding studies: NIH 3T3 and LMTK- cells had significantly fewer abrin binding sites than MEF and S-180 cells. The fate of the [125I]abrin after internalization was examined by gel electrophoresis and autoradiography. A pattern of time-dependent degradation was observed, degradation being more rapid in NIH 3T3 and S-180 cells than in LMTK- and MEF cells. We conclude that the varying responses of different cells to the toxin abrin may be due to several factors, including the relative number of abrin binding sites on the cell surface and the rate of degradation of the toxin once internalized. The results also show that the sensitivities of the cells to abrin do not necessarily correlate with their normal or neoplastic state.  相似文献   

18.
The nature of binding of abrin to Chinese hamster ovary cells was examined in relation to the ensuing intoxication of the treated cells. Approx. 20% of [125I]abrin bound to CHO cells at 37°C was found to be resistant to the addition or presence of 0.1 M lactose. The extent of lactose-resistant binding depended inversely upon the temperature of incubation. Among various proteins, lectins and sugars, only non-labeled abrin could strongly inhibit the lactose-resistant binding of [125I]abrin. Lactose-resistant binding could lead to an inhibition of cellular protein synthesis and to a loss of cell viability. Abrin molecules bound at the lactose-sensitive and lactose-resistant binding sites apparently have an equal probability of being internalized by CHO cells. Binding of approx. 3·103 abrin molecules per CHO cell was required to elicit 50% loss of cell viability regardless of whether the binding occurs in the presence or absence of lactose. The result of a cross-linking experiment suggested that a membrane protein with an Mr of about 45 000 may be responsible for the lactose-resistant binding of abrin.  相似文献   

19.
A. Sakai  Y. Shimizu    F. Hishinuma 《Genetics》1988,119(3):499-506
We have isolated mutants responsible for an oversecretion phenotype in Saccharomyces cerevisiae, using a promoter of SUC2 and the gene coding for alpha-amylase from mouse as a marker of secretion. These mutations defined two complementation groups, designated as ose1 (over secretion) and rgr1 (resistant to glucose repression). The ose1 mutant produced an oversecretion of amylase by 12- to 15-fold under derepressing conditions; however, the amylase mRNA was present at nearly the same amount as it was in the parent cells. No expression of the amylase gene was detected under repressing conditions. The rgr1 mutant oversecreted amylase by 11- to 13-fold under repressing conditions by 15- to 18-fold under derepressing conditions. The rgr1 mutant showed pleiotropic effects on the following cellular functions: (1) resistance to glucose repression, (2) temperature-sensitive lethality, (3) sporulation deficieny in homozygous diploid cells, and (4) abnormal cell morphology. The rgr1 mutation was not allelic with ssn6 and cyc9, and failed to suppress snf1.  相似文献   

20.
Cell walls of Staphylococcus aureus R9/80 resistant to gramicidin S and actinomycin D were investigated. The strain was isolated after passages of a previously isolated strain of S. aureus with resistance to gramicidin and definite changes in the cell walls, a medium with increasing concentrations of actinomycin being used for the passages. The data on the study of the cell walls of the strain with the double resistance were compared with the results of the investigation of the cell walls of the strain susceptible to gramicidin, the gramicidin resistant strain (initial for strain R9/80) and the actinomycin adapted strain that also showed changes in the cell walls. The cell walls of the resistant strains had no significant changes in the peptidoglycane and glucosamine levels, as well as in the peptidoglycane amino acid composition. Teichoic acids of all the strains had different levels of substitution of ribite by D-alanine (a factor influencing the negative charge of teichoic acids and the wall at large). It was noted that all the strains resistant to the tested antibiotics had lower levels of teichoic acids in the cell walls. The resistant cells showed some increase of the lipid component in the walls: from 1.6% in the susceptible strain to 2.1-2.9% in the resistant cells. The main trend of the changes in the resistance development was revealed to be the thickening of the cell wall and its consolidation. The development of resistance to gramicidin, actinomycin and to both the antibiotics provoked respectively a 2.4-, 4- and 5.4-fold increase of the content of the main cell component. i.e. peptidoglycane in the cell biomass. The barrier role of the cell walls in the resistant strains and their ability to bind the antibiotic is discussed.  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