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.
An aqueous acetone extract obtained from the pericarps of Mallotus japonicus (MJE) was observed to inhibit prostaglandin (PG) E(2) production in a lipopolysaccharide (LPS)-activated murine macrophage-like cell line, RAW 264.7. Six phloroglucinol derivatives isolated from MJE exhibited inhibitory activity against PGE(2) production. Among these phloroglucinol derivatives, isomallotochromanol showed the strongest inhibitory activity, with an IC(50) of 1.0 microM. MJE and its phloroglucinol derivatives did not effect the enzyme activity of either prostaglandin endoperoxide synthase (PGHS)-1 or PGHS-2. However, induction of PGHS-2 in LPS-activated macrophages was inhibited by MJE and its phloroglucinol derivatives, whereas the level of PGHS-1 protein was not affected. Moreover, RT-PCR analysis showed that MJE and its phloroglucinol derivatives significantly suppressed PGHS-2 mRNA expression. Therefore, the observed inhibition of PGHS-2 induction by MJE and its phloroglucinol derivatives was likely due to a suppression of PGHS-2 mRNA expression. These results suggest that MJE and its phloroglucinol derivatives have the pharmacological ability to suppress PGE(2) production by activated macrophages.  相似文献   

2.
An aqueous acetone extract of the pericarps of Mallotus japonicus (MJE) inhibited nitric oxide (NO) production by a murine macrophage-like cell line, RAW 264.7, which was activated by lipopolysaccharide (LPS) and interferon-gamma (IFN-gamma). Seven phloroglucinol derivatives isolated from MJE exhibited inhibitory activity against NO production. Among these phloroglucinol derivatives, isomallotochromanol exhibited strong inhibitory activity toward NO production, exhibiting an IC(50) of 10.7 microM. MJE and the phloroglucinol derivatives significantly reduced both the induction of inducible nitric oxide synthase (iNOS) protein and iNOS mRNA expression. NO production by macrophages preactivated with LPS and IFN-gamma for 16 h was also inhibited by MJE and the phloroglucinol derivatives. Furthermore, MJE and the derivatives directly affected the conversion of L-[(14)C]arginine to L-[(14)C]citrulline by the cell extract. These results suggest that MJE and the phloroglucinol derivatives have the pharmacological ability to suppress NO production by activated macrophages. They inhibited NO production by two mechanisms: reduction of iNOS protein induction and inhibition of enzyme activity.  相似文献   

3.
In the literature, biological tyrosine nitrations have been reported to depend not only on peroxynitrite but also on nitrite/hydrogen peroxide linked to catalysis by myeloperoxidase. In endotoxin-stimulated RAW 264.7 macrophages, we have detected a major nitrotyrosine positive protein band around 72 kDa and identified it as prostaglandin endoperoxide synthase-2 (PGHS-2). Isolated PGHS-2 in absence of its substrate arachidonate was not only tyrosine-nitrated with peroxynitrite, but also with nitrite/hydrogen peroxide in complete absence of myeloperoxidase. Our data favor an autocatalytic activation of nitrite by PGHS-2 with a subsequent nitration of the essential tyrosine residue in the cyclooxygenase domain. Under inflammatory conditions, nitrite formed via NO-synthase-2 may therefore act as an endogenous regulator for PGHS-2 in stimulated macrophages. Nitration of PGHS-2 by the autocatalytic activation of nitrite further depends on the intracellular concentration of arachidonate since arachidonate reacted competitively with nitrite and could prevent PGHS-2 from nitration when excessively present.  相似文献   

4.
Nystatin is known to deplete lipid rafts from mammalian cell membranes. Lipid rafts have been reported to be necessary for lipopolysaccharide signaling. In this study, it was unexpectedly found that lipopolysaccharide-induced nitric oxide production was not inhibited, but rather increased in the presence of a non-cytotoxic concentration of nystatin. Surprisingly, treatment with nystatin induced only NO production and iNOS expression in RAW264.7 cells. At the concentration used, no changes in the expression of GM1 ganglioside, a lipid raft marker on RAW264.7 cells, was seen. From studies using several kinds of inhibitors for signaling molecules, nystatin-induced NO production seems to occur via the iκB/NF-κB and the PI3 K/Akt pathway. Furthermore, because nystatin is known to activate the Na-K pump, we examined whether the Na-K pump inhibitor amiloride suppresses nystatin-induced NO production. It was found that amiloride significantly inhibited nystatin-induced NO production. The results suggest that a moderate concentration of nystatin induces NO production by Na-pump activation through the PI3 kinase/Akt/NF-κB pathway without affecting the condition of lipid rafts.  相似文献   

5.
The rate of arachidonic acid metabolism by prostaglandin (PG) endoperoxide synthetase by the rabbit renal cortex (approximately 200 pmol/mg of protein/min) is very slow compared to medulla (approximately 2000 pmol/mg/min). However, by using PGH2 as a substrate and limiting reaction times, we were able to directly measure the endoperoxide-dependent PGE2 isomerase and found that both the cortex and medulla possess high levels of this activity (approximately pmol/mg/min). The PG endoperoxide E2 isomerase is dependent on reduced glutathione, but not cysteine, and is inactivated by p-hydroxymercuribenzoate or boiling. Thus, the renal medulla appears to exhibit an efficient coupling of cyclooxygenase and PG endoperoxide E2 isomerase, whereas the cortex has a vast excess of endoperoxide-dependent enzyme.  相似文献   

6.
7.
Prostaglandin E2 (PGE2) is the major cyclooxygenase metabolite in macrophages with complex proinflammatory and immunoregulatory properties. In the present study, we have compared the modulatory role of PGE2/cAMP-dependent signaling on induced nitric oxide (NO) production in two murine macrophages, J774 and RAW 264.7. With no effect on NO release by itself, PGE2 co-addition with lipopolysaccharide (LPS) resulted in a concentration-dependent enhancement in NO release and inducible NO synthase induction in J774, but not in RAW 264.7, macrophages. The potentiation effect of PGE2 in J774 cells was still seen when applied within 9 h after LPS treatment. Whereas RAW 264.7 macrophages release PGE2 with greater extent than J774 macrophages in response to LPS, indomethacin and NS-398, upon abolishing LPS-induced PGE2 release, caused a more obvious inhibition of NO release from J774 than RAW 264.7 cells. Thus, we suggest a higher positive modulatory role of PGE2--either endogenous or exogenous--on NO formation in J774 cells. Supporting these findings, exogenous PGE2 triggers cAMP formation in J774 cells with higher potency and efficacy. Of interest, dBcAMP also elicits higher sensitivity in potentiating NO release in J774 cells. We conclude that the opposite effect of PGE2/cAMP signaling on macrophage NO induction depends on its signaling efficacy and might be associated with the difference in endogenous PGE2 levels.  相似文献   

8.
Current evidence suggests that two forms of prostaglandin (PG) E synthase (PGES), cytosolic PGES and membrane-bound PGES (mPGES) -1, preferentially lie downstream of cyclooxygenase (COX) -1 and -2, respectively, in the PGE2 biosynthetic pathway. In this study, we examined the expression and functional aspects of the third PGES enzyme, mPGES-2, in mammalian cells and tissues. mPGES-2 was synthesized as a Golgi membrane-associated protein, and spontaneous cleavage of the N-terminal hydrophobic domain led to the formation of a truncated mature protein that was distributed in the cytosol with a trend to be enriched in the perinuclear region. In several cell lines, mPGES-2 promoted PGE2 production via both COX-1 and COX-2 in the immediate and delayed responses with modest COX-2 preference. In contrast to the marked inducibility of mPGES-1, mPGES-2 was constitutively expressed in various cells and tissues and was not increased appreciably during tissue inflammation or damage. Interestingly, a considerable elevation of mPGES-2 expression was observed in human colorectal cancer. Collectively, mPGES-2 is a unique PGES that can be coupled with both COXs and may play a role in the production of the PGE2 involved in both tissue homeostasis and disease.  相似文献   

9.
10.
Sosroseno W  Bird PS  Seymour GJ 《Anaerobe》2011,17(5):246-251
Nitric oxide (NO) may play a crucial role in the pathogenesis of periodontal disease and, hence, the aim of the present study was to test the hypothesis that Aggregatibacter actinomycetemcomitans surface-associated material (SAM) stimulates inducible nitric oxide synthase (iNOS) activity and NO production by the murine macrophage cell line RAW264.7. Cells were stimulated with untreated or heat-treated A. actinomycetemcomitans SAM and with or without pre-treatment with l-N6-(1-Iminoethyl)-lysine (L-NIL) (an iNOS inhibitor), polymyxin B, interferon-gamma (IFN-γ) and Interleukin-4 (IL-4), IL-10, genistein [a protein tyrosine kinase (PTK) inhibitor], bisindolylmaleimide [a protein kinase C (PKC) inhibitor], bromophenacyl bromide (BPB) [a phospholipase A2 (PLA2) inhibitor] or wortmannin [phosphatidylinositol 3-kinase (PI-3K) inhibitor]. The iNOS activity and nitrite production in the cell cultures were determined. Untreated but not heat-treated A. actinomycetemcomitans SAM-stimulated both iNOS activity and nitrite production in RAW264.7 cells. l-NIL, IL-4, IL-10, genistein, bisindolylmaleimide, or BPB, suppressed but IFN-γ enhanced both iNOS activity and nitrite production by A. actinomycetemcomitans SAM-stimulated cells. Wortmannin and polymyxin B failed to alter both iNOS activity or nitrite production by A. actinomycetemcomitans SAM treated cells. Therefore, the present study suggests that a heat-sensitive protein constituent(s) of A. actinomycetemcomitans SAM stimulates both iNOS activity and nitrite production by RAW264.7 cells in a cytokine, PTK, PKC, and PLA2 but not PI-3K-dependent fashion.  相似文献   

11.
INTRODUCTION: Prostaglandins (PGs) can act on both hematopoietic and osteoblastic lineages to enhance osteoclast formation. METHODS: We examined PGE2 stimulated osteoclastogenesis in RAW 264.7 cells and the role of endogenous PGE2 in lipopolysaccharide (LPS) stimulated osteoclastogenesis. RESULTS: RANKL (1-100 ng/ml) increased formation of osteoclasts, defined as tartrate resistant acid phosphatase multinucleated cells, with peak effects at 30 ng/ml. Addition of PGE2 (0.01-1.0 microM) to RANKL (30 ng/ml) dose dependently increased osteoclast number 30-150%. Use of NS-398 (0.1 microM) or indomethacin (Indo, 1.0 micro M) to block endogenous PG synthesis had little effect on the response to RANKL alone but significantly decreased the response to PGE2. Addition of LPS (100 ng/ml) to RANKL increased osteoclast number 50%, and this response was significantly decreased by NS-398 and Indo. RANKL and PGE2 produced small, additive increases in COX-2 mRNA levels, while LPS produced a larger increase. PG release into the medium was not increased by RANKL and PGE2 but markedly increased by LPS. CONCLUSION: We conclude that RANKL stimulated osteoclastogenesis can be enhanced by PGE2 and LPS though direct effects on the hematopoietic cell lineage and that these effects may be mediated in part by induction of COX-2 and enhanced intracellular PG production.  相似文献   

12.
This study examined the effect of prostaglandin E2 (PGE2) produced by microsomal prostaglandin E synthase-1 (mPGES-1) on circadian rhythm. Using wild-type mice (WT) and mPGES-1 knockout mice (mPGES-1−/−), I recorded and automatically analyzed the natural behavior of mice in home cages for 24 h and measured brain levels of PGE2. The switch to wakefulness was not smooth, and sleepiness and the total duration of sleep were significantly longer in the mPGES-1−/− mice. Moreover, the basal concentration of PGE2 was significantly lower in the mPGES-1−/− mice. These findings suggest that PGE2 produced by mPGES-1 regulates the onset of wakefulness and the maintenance of circadian rhythm.  相似文献   

13.
Combinatorial chemopreventive strategies, in contrast to those with individual agents, show potential in terms of potentially lower toxicity and higher efficacy. In this study, we combined several agents and examined their suppressive effects on the combined lipopolysaccharide (LPS)- and interferon(IFN)-gamma-induced formation of proinflammatory mediators, including prostaglandin (PG) E2 and tumor necrosis factor (TNF)-alpha, in RAW264.7 murine macrophages. The combinatorial effects of indomethacin/genistein (GEN) and aspirin/GEN were found to be synergistic for PGE2 suppression, while the nimesulide/GEN combination was antagonistic. Further, while (-)-epigallocatechin gallate (EGCG) alone increased LPS/IFM-gamma-induced production of PGE2 and TNF-alpha as well as cyclooxygenase-2 expression, the EGCG/GEN combination markedly suppressed these parameters. Our results suggest that certain chemopreventive agents act complexly and that, when used in combination, they affect the intracellular signaling pathways of the paired agents to exert additive, synergistic, or antagonistic effects.  相似文献   

14.
Increased expression of prostaglandin endoperoxide H synthase-2 (PGHS-2) has been implicated in pathological conditions such as inflammatory bowel diseases and colon cancer. Recently, it has been demonstrated that inducible nitric oxide synthase (NOS II) expression and nitric oxide (NO) production are up-regulated in these diseases as well. However, the apparent link between PGHS-2 and NOS II has not been thoroughly investigated in nontransformed and nontumorigenic colonic epithelial cells. In the present study, we examined the concomitant expression of PGHS-2 and NOS II as well as the production of prostaglandin E2 (PGE2) and NO in conditionally immortalized mouse colonic epithelial cells, namely YAMC (Apc(+/+)). We found that the induction of PGHS-2 and generation of PGE2 in these cells by IFN-gamma and lipopolysaccharide (LPS) were greatly reduced by two selective NOS II inhibitors, L-NIL and SMT. To ascertain the effect of NO on PGHS-2 overexpression, we tested NO-releasing compounds, NOR-1 and SNAP, and found that they caused PGHS-2 expression and PGE2 production. This effect was abolished by hemoglobin, a NO scavenger. Using electrophoretic mobility shift assays, we found that both NOR-1 and SNAP caused beta-catenin/LEF-1 DNA complex formation. Super-shift by anti-beta-catenin antibody confirmed the presence of beta-catenin in the complex. Cell fractionation studies indicated that NO donors caused an increase in free soluble cytoplasmic beta-catenin. This is further corroborated by the immunocytochemistry data showing the redistribution of beta-catenin from the predominantly membrane localization into the cytoplasm and nucleus after treatment with NO donors. To further explore the possible connection between PGHS-2 expression and beta-catenin/LEF-1 DNA complex formation, we studied IMCE (Apc(Min/+)) cells, a sister cell line of YAMC with similar genetic background but differing in Apc genotype and, consequently, their beta-catenin levels. We found that IMCE cells, in comparison with YAMC cells, had markedly higher beta-catenin/LEF-1 DNA complex formation under both resting conditions as well as after induction with NO. In parallel fashion, IMCE cells expressed significantly higher levels of PGHS-2 mRNA and protein, and generated more PGE2. Overall, this study suggests that NO may be involved in PGHS-2 overexpression in conditionally immortalized mouse colonic epithelial cells. Although the molecular mechanism of the link is still under investigation, this effect of NO appears directly or indirectly to be a result of the increase in free soluble beta-catenin and the formation of nuclear beta-catenin/LEF-1 DNA complex.  相似文献   

15.
Kim HW  Kim JH  An HS  Park KK  Kim BK  Park T 《Life sciences》2003,73(19):2477-2489
The role of myo-inositol in the regulation of taurine transport in activated murine macrophage cell line, RAW 264.7, was studied. Challenge of RAW 264.7 murine macrophages for 24 hr with phorbol ester 12-myristate 13-acetate (PMA) (10 ng/ml), a PKC activator, resulted in a 62% decrease in taurine transport activity. Among the various monosaccharides (1 mM) tested in the presence of PMA, myo-inositol was most effective in restoring the PMA-induced down-regulation of taurine transport in murine macrophages (82% increase compared to the value for cells treated with PMA Alone, p < 0.01). The protective role of myo-inositol against stress-induced down-regulation of taurine transport by macrophages was further investigated in conditions mimicking bacterial infection, inflammation, and immune-suppressed circumstances. A challenge of murine macrophages with lipopolysaccharide (LPS) (0.1 and 10 microg/ml) resulted in a 60% decrease in taurine transport activity compared to the value for untreated control cells (p < 0.01). When cells were co-treated with myo-inositol (100 nM approximately 10 mM) in the presence of LPS for 24 hrs, taurine transport activity increased in a dose-dependent manner compared to the value for cells treated with LPS only. Taurine transport activity in cells treated with LPS (10 microg/ml) plus interferon-gamma (IFN-gamma) (150 unit/ml) for 24 hrs was 13% of the value for untreated control cells (p < 0.01). Again, this inflammation-induced down-regulation of taurine transport activity was completely antagonized with co-administration of 100 nM or higher levels of myo-inositol in the culture medium. Similarly, myo-inositol effectively restored the taurine transport activity suppressed by cyclosporin A (0.5 and 50 nM) in murine macrophages (p < 0.01). From these results, myo-inositol appears to be a common accelerator of taurine transport by murine macrophages in diverse conditions of down-regulated taurine transport.  相似文献   

16.
Microsomal prostaglandin E synthase (mPGES)-1, which is dramatically induced in macrophages by inflammatory stimuli such as lipopolysaccharide (LPS), catalyzes the conversion of cyclooxygenase-2 (COX-2) reaction product prostaglandin H(2) (PGH(2)) into prostaglandin E(2) (PGE(2)). The mPGES-1-derived PGE(2) is thought to help regulate inflammatory responses. On the other hand, excess PGE(2) derived from mPGES-1 contributes to the development of inflammatory diseases such as arthritis and inflammatory pain. Here, we examined the effects of liver X receptor (LXR) ligands on LPS-induced mPGES-1 expression in murine peritoneal macrophages. The LXR ligands 22(R)-hydroxycholesterol (22R-HC) and T0901317 reduced LPS-induced expression of mPGES-1 mRNA and mPGES-1 protein as well as that of COX-2 protein. However, LXR ligands did not influence the expression of microsomal PGES-2 (mPGES-2) or cytosolic PGES (cPGES) protein. Consequently, LXR ligands suppressed the production of PGE(2) in macrophages. These results suggest that LXR ligands diminish PGE(2) production by inhibiting the LPS-induced gene expression of the COX-2-mPGES-1 axis in LPS-activated macrophages.  相似文献   

17.
To better define the role of the various prostanoid synthases in the adjuvant-induced arthritis (AIA) model, we have determined the temporal expression of the inducible PGE synthase (mPGES-1), mPGES-2, the cytosolic PGES (cPGES/p23), and prostacyclin synthase, and compared with that of cyclooxygenase-1 (COX-1) and COX-2. The profile of induction of mPGES-1 (50- to 80-fold) in the primary paw was similar to that of COX-2 by both RNA and protein analysis. Quantitative PCR analysis indicated that induction of mPGES-1 at day 15 was within 2-fold that of COX-2. Increased PGES activity was measurable in membrane preparations of inflamed paws, and the activity was inhibitable by MK-886 to >or=90% with a potency similar to that of recombinant rat mPGES-1 (IC(50) = 2.4 microM). The RNA of the newly described mPGES-2 decreased by 2- to 3-fold in primary paws between days 1 and 15 postadjuvant. The cPGES/p23 and COX-1 were induced during AIA, but at much lower levels (2- to 6-fold) than mPGES-1, with the peak of cPGES/p23 expression occurring later than that of COX-2 and PGE(2) production. Prostacyclin (measured as 6-keto-PGF(1alpha)) was transiently elevated on day 1, and prostacyclin synthase was down-regulated at the RNA level after day 3, suggesting a diminished role of prostacyclin during the maintenance of chronic inflammation in the rat AIA. These results show that mPGES-1 is up-regulated throughout the development of AIA and suggest that it plays a major role in the elevated production of PGE(2) in this model.  相似文献   

18.
We have previously demonstrated that dietary histamine is accumulated in the spleens of L-histidine decarboxylase (HDC)-deficient mice, which lack endogenous histamine synthesis. To characterize the clearance system for dietary histamine in mice, we investigated the cell type and mechanism responsible for histamine uptake in the spleens of HDC-deficient mice. Immunohistochemical analyses using an antihistamine antibody indicated that a portion of the CD14+ cells in the spleen is involved in histamine storage. Peritoneal macrophages obtained from Balb/c mice and a mouse macrophage cell line, RAW264.7, had potential for histamine uptake, which was characterized by a low affinity and high capacity for histamine. The histamine uptake by RAW264.7 cells was observed at physiological temperature and was potently inhibited by pyrilamine, chlorpromazine, quinidine, and chloroquine, moderately inhibited by N-methylhistamine, dopamine, and serotonin, and not affected by tetraethylammonium and 1-methyl-4-phenylpyridinium. Intracellular histamine was not metabolized in RAW264.7 cells and was released at physiological temperature in the absence of extracellular histamine. These results suggest that histamine uptake by macrophages may be involved in the clearance of histamine in the local histamine-enriched environment. cation transporter; chlorpromazine; pyrilamine; quinidine  相似文献   

19.
Prostaglandins (PGs) have numerous cardiovascular and inflammatory effects. Cyclooxygenase (COX), which exists as COX-1 and COX-2 isoforms, is the first enzyme in the pathway in which arachidonic acid is converted to PGs. Prostaglandin E2 (PGE2) exerts a variety of biological activities for the maintenance of local homeostasis in the body. Elucidation of PGE2 involvement in the signalling molecules such as COX could lead to potential therapeutic interventions. Here, we have investigated the effects of PGE2 on the induction of COX-2 in human umbilical vein endothelial cells (HUVEC) treated with interleukin-1beta (IL-1beta 1 ng/ml). COX activity was measured by the production of 6-keto-PGF1alpha, PGE2, PGF2alpha and thromboxane B2 (TXB2) in the presence of exogenous arachidonic acids (10 microM for 10 min) using enzyme immunoassay (EIA). COX-1 and COX-2 protein was measured by immunoblotting using specific antibody. Untreated HUVEC contained only COX-1 protein while IL-1beta treated HUVEC contained COX-1 and COX-2 protein. PGE2 (3 microM for 24h) did not affect on COX activity and protein in untreated HUVEC. Interestingly, PGE2 (3 microM for 24h) can inhibit COX-2 protein, but not COX-1 protein, expressed in HUVEC treated with IL-1beta. This inhibition was reversed by coincubation with forskolin (100 microM). The increased COX activity in HUVEC treated with IL-1beta was also inhibited by PGE2 (0.03, 0.3 and 3 microM for 24h) in a dose-dependent manner. Similarly, forskolin (10, 50 or 100 microM) can also reverse the inhibition of PGE2 on increased COX activity in IL-1beta treated HUVEC. The results suggested that (i) PGE2 can initiate negative feedback regulation in the induction of COX-2 elicited by IL-1beta in endothelial cells, (ii) the inhibition of PGE2 on COX-2 protein and activity in IL-1beta treated HUVEC is mediated by cAMP and (iii) the therapeutic use of PGE2 in the condition which COX-2 has been involved may have different roles.  相似文献   

20.
Peroxiredoxin (PRX), a scavenger of H2O2 and alkyl hydroperoxides in living organisms, protects cells from oxidative stress. Contrary to its known anti‐oxidant roles, the involvement of PRX‐1 in the regulation of lipopolysaccharide (LPS) signaling is poorly understood, possible immunological functions of PRX‐1 having been uncovered only recently. In the present study, it was discovered that the PRX‐1 deficient macrophage like cell line (RAW264.7) has anti‐inflammatory activity when stimulated by LPS. Treatment with LPS for 3 hrs resulted in increased gene expression of an anti‐inflammatory cytokine, interleukin‐10 (IL‐10), in PRX‐1 knock down RAW264.7 cells. Gene expression of pro‐inflammatory cytokines IL‐1β and tumor necrosis factor‐ α (TNF‐α) did not show notable changes under the same conditions. However, production of these cytokines significantly decreased in PRX‐1 knock down RAW264.7 cells with 12 hrs of stimulation. Production of IL‐10 was also increased in PRX‐1 knock down RAW264.7 cells with 12 hrs of stimulation. We predicted that higher concentrations of IL‐10 would result in decreased expression of IL‐1β and TNF‐α in PRX‐1 knock‐down cells. This was confirmed by blocking IL‐10, which reestablished IL‐1β and TNF‐α secretion. We also observed that increased concentrations of IL‐10 do not affect the NF‐κB pathway. Interestingly, STAT3 phosphorylation by LPS stimulation was significantly increased in PRX‐1 knockdown RAW264.7 cells. Up‐regulation of IL‐10 in PRX‐1 knockdown cells and the resulting downregulation of proinflammatory cytokine production seem to involve the STAT3 pathway in macrophages. Thus, down‐regulation of PRX‐1 may contribute to the suppression of adverse effects caused by excessive activation of macrophages through affecting the STAT3 signaling pathway.  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