Methanol extracts of Euphorbia cooperi inhibit the production of inflammatory mediators by inhibiting the activation of c‑Jun N‑terminal kinase and p38 in murine macrophages

  • Authors:
    • Young‑Chang Cho
    • In‑Seon Lee
    • Huiyun Seo
    • Anna Ju
    • Deokkyu  Youn
    • Younghyun Kim
    • Jaehee Choun
    • Sayeon Cho
  • View Affiliations

  • Published online on: September 11, 2014     https://doi.org/10.3892/mmr.2014.2560
  • Pages: 2663-2668
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Numerous Euphorbiaceae plants have been used for the treatment of diseases, including liver diseases, asthma and rheumatism. The present study evaluated the effect of methanol extracts from Euphorbia cooperi (MEC), a member of the Euphorbiaceae plant family, on the production of inflammatory cytokines interleukin (IL)‑6 and tumor necrosis factor (TNF)‑α, nitric oxide (NO) as well as the activation of mitogen‑activated protein kinase and nuclear factor (NF)‑κB signaling. Non‑cytotoxic concentrations of MEC significantly reduced the production of NO and IL‑6, but not TNF‑α, in lipopolysaccharide (LPS)‑stimulated RAW 264.7 macrophages. The decreased production of NO by MEC was due to alleviated expression of inducible NO synthase. Reporter assays with cells treated with MEC demonstrated reduced activator protein‑1 (AP-1) activity, while NF‑κB activity was not reduced. Furthermore, the phosphorylation levels of c‑Jun N‑terminal kinase (JNK) and p38 were suppressed by MEC while phosphorylation levels of inhibitor of κB were not reduced by MEC, suggesting that MEC‑mediated inactivation of JNK and p38 is the underlying regulatory mechanism for inflammatory mediators in LPS‑stimulated RAW 264.7 macrophages.

Introduction

Macrophages are the primary effector cells of the immune system that protect against microbial infection. Upon stimulation with pathogen-derived molecules, including lipopolysaccharide (LPS) and flagellin, macrophages secrete a variety of inflammatory mediators and cytokines (1,2). These secreted proteins bind tightly to toll-like receptors (TLRs) and stimulate the formation of signaling complexes resulting in a rapid defensive response (36). The tight regulation of host immune signaling ensures that the resulting immune response is appropriate for the continuously changing microenvironment, as well as for maintenance of immunological balance. However, inappropriate or prolonged activation of the immune system is largely responsible for the pathology of acute and chronic inflammatory conditions, including septic shock and chronic inflammatory conditions, such as rheumatoid arthritis, inflammatory bowel disease and chronic obstructive pulmonary disease (1,7). In this regard, the secretion of pro-inflammatory mediators by activated macrophages, including cytokines, growth factors, hydrolytic enzymes, bioactive lipids, reactive oxygen intermediates and nitric oxide (NO), are involved in the pathogenesis of tissue injury (1,3). Therefore, identification of agents that can regulate the production of pro-inflammatory mediators is considered an effective strategy for developing therapeutic agents to treat severe inflammation.

Activated macrophages transcriptionally express inducible NO synthase (iNOS) in response to various pro-inflammatory cytokines and bacterial LPS, resulting in the production of NO via the oxidative deamination of L-arginine at sites of inflammation (8,9). NO modulates a variety of biological processes, including inflammation and carcinogenesis (10,11). Tumor necrosis factor-α (TNF-α) and interleukin (IL)-6 are potent pro-inflammatory cytokines that are important for stimulating the secretion of additional inflammatory cytokines. Therefore, inhibiting the excessive production of these mediators in macrophages through the inhibition of mRNA and protein expression of iNOS, TNF-α and IL-6 may be a viable strategy for the development of novel anti-inflammatory agents.

Several attempts have been made to develop a new generation of anti-inflammatory agents from natural compounds, as natural compounds are known to have fewer side effects (12). The latexes of Euphorbia plants were traditionally considered as toxic substances due to their irritancy on mucus membranes, including the nose and mouth, in humans (13). Although Euphorbia plants have toxic effects in humans, the extracts of Euphorbia plants are well known to inhibit excessive inflammation and are used in Chinese medicine (14,15). Suarez et al demonstrated that intraperitoneal administration of an aqueous extract of Croton malambo (Euphorbiaceae) resulted in a significant anti-inflammatory effect in a rat model of edema (16). Furthermore, Sangre de Drago (dragon’s blood) from Croton lechleri (Euphorbiaceae) inhibits inflammation in vitro and in vivo (17,18). For example, treatment with Sangre de Drago significantly decreased intracellular generation of reactive oxygen species in several cell lines and alleviated paw edema in rats (18). However, the anti-inflammatory effects of methanol extracts of Euphorbia cooperi (MEC) remain to be elucidated.

In the present study, the anti-inflammatory effect of MEC in LPS-stimulated RAW 264.7 macrophages and its underlying mechanisms were investigated to evaluate the therapeutic potential of MEC in the treatment of abnormal inflammation.

Materials and methods

Cell culture and reagents

The RAW 264.7 macrophages, a mouse monocytic cell line, were cultured in Dulbecco’s modified Eagle’s medium supplemented with 10% fetal bovine serum, 50 U/ml penicillin and 50 μg/ml streptomycin (Gibco-BRL, Grand Island, NY, USA) at 37°C in a 5% CO2 humidified air atmosphere. Rabbit polyclonal anti-inhibitors of κB (IκB) and mouse anti-tubulin antibodies were purchased from Santa Cruz Biotechnology, Inc. (Santa Cruz, CA, USA). Rabbit polyclonal anti-inducible iNOS, rabbit polyclonal anti-phospho IκB, rabbit polyclonal anti-phospho p38 mitogen-activated protein kinase (MAPK), rabbit polyclonal anti-p38, mouse monoclonal anti-phospho extracellular signal-regulated kinase (ERK), rabbit polyclonal anti-ERK, rabbit polyclonal anti-phospho c-Jun N-terminal kinase (JNK) and mouse monoclonal anti-JNK were purchased from Cell Signaling Technology Inc. (Danvers, MA, USA). MEC was purchased from the International Biological Material Research Center (Daejeon, Korea). The above compounds were dissolved in dimethyl sulfoxide (DMSO; Sigma-Aldrich, St. Louis, MO, USA) and added directly to the culture media. The final concentrations of DMSO never exceeded 0.1%, which did not affect the assay systems.

3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay

The RAW 264.7 macrophages were incubated with MEC and LPS for 24 h. Following incubation, MTT (0.5 mg/ml) was added for 3 h at 37°C and the supernatants were carefully removed. The crystals of viable cells were dissolved in DMSO and absorbance was measured at 595 nm using a Synergy microplate reader (BioTek Instruments Inc., Winooski, VT, USA).

Nitrite assay

The RAW 264.7 macrophages were incubated with MEC and LPS for 24 h. Following incubation, the levels of NO synthesis were determined by assaying the culture supernatants for nitrite, the stable reaction product of NO, with molecular oxygen, using the Griess reagent (1% sulfanilamide, 0.1% N-1-naphthylenediamine dihydrochloride and 2.5% phosphoric acid). The absorbance was measured at 540 nm using a Synergy microplate reader after 10 min incubation.

Enzyme-linked immunosorbent assay (ELISA)

The RAW 264.7 macrophages were stimulated with LPS and MEC for 24 h. Following stimulation, the supernatants were obtained and the quantities of TNF-α and IL-6 in the culture supernatants were determined using sandwich ELISA, which used monoclonal antibodies specific to each mediator. Prior to the application of samples, the plate was pre-coated with coating antibody in the supplied buffer. Following overnight incubation at 4°C, the plate was washed and assay diluents (1X) were treated for 1 h. The samples were then loaded into each well and incubated for 2 h at room temperature. They were then treated with biotinylated secondary antibody and horseradish peroxidase-streptavidin solutions for 1 h and 30 min, respectively and substrate solution was added to the washed plate. After 10 min incubation in the dark, 1N H3PO4 treatment was applied and the optical density of the individual wells was determined at 450 nm using a Synergy microplate reader.

Reverse transcription polymerase chain reaction (RT-PCR)

Total RNA was prepared from the cells and reverse-transcribed into complementary DNA (cDNA), following which PCR amplification of the cDNA was performed. The sequences of PCR primers used in the present study were as follows: mouse iNOS, forward 5′-GCA TGG AAC AGT ATA AGG CAA ACA-3′ and reverse 5′-GTT TCT GGT CGA TGT CAT GAG CAA-3′; TNF-α, forward 5′-GTG CCA GCC GAT GGG TTG TAC C-3′ and reverse 5′-AGG CCC ACA GTC CAG GTC ACT G3′; IL-6, forward 5′-TCT TGG GAC TGA TGC TGG TGA C-3′ and reverse 5′-CAT AAC GCA CTA GGT TTG CCG A-3′ and GAPDH, forward 5′-GTC TTC ACC ACC ATG GAG AAG G-3′ and reverse 5′-CCT GCT TCA CCA CCT TCT TGC C-3′. The PCR was run for 20–25 cycles of 94°C (30 sec), 60°C (30 sec) and 72°C (30 sec) using a Bioer’s thermal cycler (Bioer Technology Co., Hangzhou, China). Following amplification, the RT-PCR products (10 μl) were separated in 1.5% (w/v) agarose gels and stained with ethidium bromide.

Transient transfection and luciferase assay

The nuclear factor-kappa B (NF-κB) and activator protein-1 (AP-1) promoters containing the luciferase gene were purchased from Agilent Technologies (Santa Clara, CA, USA). HEK 293 cells were transiently transfected using polyethyleneimine according to the manufacturer’s instructions (Polysciences Inc., Warrington, PA, USA) and transfected cells were stimulated with phorbol 12-myristate 13-acetate (PMA) in the presence or absence of MEC for 24 h. The cells were harvested and the luciferase activities were assayed according to the manufacturer’s instructions (Promega Corporation, Madison, WI, USA).

Preparation of total cell lysates

LPS-stimulated RAW 264.7 cells were treated with MEC for the indicated time periods (15 min and 24 h) and washed with ice-cold phosphate-buffered saline. The cells were lysed in lysis buffer containing 0.5% NP-40, 0.5% Triton X-100, 150 mM sodium chloride, 20 mM trisaminomethane-hydrochloride (Tris-HCl; pH 8.0), 1 mM ethylenediaminetetraacetic acid, 1% glycerol, 1 mM phenylmethylsulfonyl fluoride and 1 μg/ml aprotinin, collected into microtubes and then centrifuged at 15,500 × g for 30 min at 4°C. The supernatants were prepared in new microtubes.

Western blot analysis

Protein concentration was measured using the Bradford method. Aliquots of the cell lysates were separated on a 10% sodium dodecyl sulfate-polyacrylamide gel in a Mini-Protein II gel apparatus (Bio-Rad, Richmond, CA, USA) and transferred onto nitrocellulose membranes (GE Healthcare, Milwaukee, WI, USA) with transfer buffer [(192 mM glycine, 25 mM Tris-HCl; pH 8.8 and 20% MeOH (v/v)]. Following inhibition of the non-specific sites with 5% bovine serum albumin solution, the membrane was incubated overnight at 4°C with the primary antibodies (1:1,000). Each membrane was further incubated for 1 h with secondary peroxidase-conjugated goat polyclonal immunoglobulin G (IgG; 1:5,000). The target proteins were detected using an enhanced chemiluminescence solution.

Statistical analysis

Differences between the experimental conditions were assessed using Student’s t-test. P<0.05 was considered to indicate a statistically significant difference. In all instances, the means of data from three independent experiments were analyzed.

Results

Effects of MEC on the viability of activated macrophages

To determine the maximal effective concentration of MEC that has minimal cytotoxicity, RAW 264.7 macrophages were treated with the indicated concentrations of MEC (20, 50, 100 and 200 μg/ml) for 24 h in the presence of LPS. Cell viability was determined by the ability of the cells to metabolically reduce a tetrazolium salt to a formazan dye. MEC had little effect on cell viability at doses of ≤100 μg/ml either in the absence or presence of 0.5 μg/ml LPS (Fig. 1). However, significant cytotoxicity was observed at MEC concentrations ≥100 μg/ml. These data indicated that low doses of MEC (100 μg/ml) do not affect the viability of RAW 264.7 macrophages. Therefore, concentrations ≤100 μg/ml were used in the subsequent experiments.

Effects of MEC on the production of LPS-induced iNOS and NO

To determine the anti-inflammatory effect of MEC, the release of NO was examined by treating the RAW 264.7 macrophages with MEC in the absence or presence of LPS. Culture supernatants were collected after 24 h incubation and the quantity of nitrite accumulated in the culture media was estimated using Griess reagent as an indicator of NO release. As shown in Fig. 2A, the nitrite concentration in the media was increased markedly in the LPS-activated RAW 264.7 macrophages compared with the unstimulated cells. When the RAW 264.7 macrophages were treated with various concentrations of MEC, the levels of LPS-stimulated nitrite production decreased significantly in a dose-dependent manner (Fig. 2A). Since nitrite is a product of iNOS activation, the effects of MEC on iNOS mRNA and protein in RAW 264.7 macrophages were measured using RT-PCR and western blot analysis, respectively. As shown in Fig. 2B, treatment of RAW 264.7 macrophages with various concentrations of MEC markedly reduced the LPS-stimulated increase in the level of iNOS mRNA expression in a dose-dependent manner. Furthermore, MEC treatment reduced the LPS-induced increase of iNOS protein expression in these cells (Fig. 2C). These results indicated that MEC reduced NO production by inhibiting the expression of iNOS in LPS-activated macrophages.

Differential inhibitory effect of MEC on the production of IL-6 and TNF-α in activated macrophages

Since stimulation of macrophages with LPS consequently induces the production of pro-inflammatory cytokines, including TNF-α and IL-6, the anti-inflammatory effects of MEC on cytokine production were evaluated in LPS-stimulated macrophages. As shown in Fig. 3A and B, LPS-stimulated RAW 264.7 macrophages produced large quantities of TNF-α and IL-6. Notably, MEC treatment reduced the LPS-stimulated production of IL-6 in a dose-dependent manner but did not affect the production of TNF-α. Furthermore, the LPS-induced increase in IL-6 mRNA expression was also markedly inhibited by MEC treatment, whereas the mRNA level of TNF-α was not affected (Fig. 3C and D). These results indicated that MEC selectively regulated the production of IL-6 by inhibiting IL-6 mRNA expression.

Selective inhibition of MAPK phosphorylation by MEC in activated macrophages

To elucidate the mechanisms underlying the anti-inflammatory effect of MEC, the activities of the inflammation-associated transcription factors, NF-κB and AP-1, were measured. HEK 293 cells transiently transfected with NF-κB or AP-1 luciferase reporter constructs were treated with PMA, either in the absence or presence of MEC and luciferase activity was determined. MEC had no apparent effect on PMA-stimulated NF-κB activity (Fig. 4A), however, the constructs were markedly stimulated by PMA and the PMA-activated AP-1 transcriptional activity was inhibited by MEC in a dose-dependent manner (Fig. 4B). To investigate this further, the effects of MEC on the LPS-induced phosphorylation of IκB and MAPKs in RAW 264.7 cells were analyzed using western blot analysis. LPS treatment markedly induced the phosphorylation of IκB and MAPKs, whereas MEC inhibited the increased phosphorylation of JNK and p38 in a dose-dependent manner. By contrast, the LPS-stimulated phosphorylation of IκB and ERK was not affected by MEC treatment (Fig. 5). The results indicated that inflammatory responses activated through MAPK signaling pathways, particularly those involving JNK and p38, are sensitive to inhibition by MEC.

Discussion

The excessive production of iNOS by external stimuli is recognized as an important pathophysiological consequence in numerous inflammatory disorders (11). Aberrant expression of iNOS leads to abnormal levels of NO production (8,19,20) and several studies have indicated that increased expression of iNOS is associated with carcinogenesis and severe inflammatory diseases, including sepsis and arthritis (10,21,22). Since iNOS/NO are central to the inflammatory process, a number of studies have attempted to identify novel anti-inflammatory agents that inhibit the expression of iNOS and to elucidate the underlying mechanism. The present study demonstrated that MEC inhibited the production of NO in RAW 264.7 macrophages by inhibiting the expression of iNOS in a non-cytotoxic manner. These results suggested that MEC contains anti-inflammatory phytochemicals.

Pro-inflammatory cytokines are key mediators of apoptosis and innate immune reactions. Excessive levels of pro-inflammatory cytokines can induce tissue injury and potentiate septic shock (23,24). Therefore, agents that inhibit the production and action of pro-inflammatory cytokines may inhibit the progression of inflammatory diseases. The pro-inflammatory cytokines, TNF-α and IL-6, are major pathogenic factors for a number of inflammatory diseases, including rheumatoid arthritis, and anti-IL-6 receptor antibody is currently used as a therapeutic agent in the clinical treatment of this disease (2426). The present study demonstrated that MEC significantly inhibited the production of IL-6, but not TNF-α, in LPS-stimulated RAW 264.7 macrophages. Several studies have demonstrated that these cytokines are not simultaneously inhibited by natural compounds (27,28). One possibility for the differential regulation of IL-6 and TNF-α by MEC is that IL-6 and TNF-α possess a different promoter binding region for transcription factors. Previous studies have revealed that the signal transducer and activator of transcription (STAT) protein binding region is contained in the IL-6 promoter region, but not in the TNF-α promoter (29), and MEC may be a regulator of STAT signaling.

Several intracellular signaling pathways are associated with the increased expression of iNOS and pro-inflammatory cytokines (2,20,30). In particular, MAPKs, including p38, ERK and JNK are important in the production of various inflammatory mediators (7,30). LPS treatment of murine macrophages significantly enhances the production of inflammatory mediators via MAPK phosphorylation and stimulation of the downstream signaling pathway (2,5,6). These studies imply that inhibition of p38, ERK and JNK phosphorylation may be a potential target pathway for the alleviation of severe inflammatory states. However, several studies have suggested that MAPK signaling cascades may be differently involved in the response of anti-inflammatory compounds in macrophages (3133). In particular, a study by Watters et al demonstrated that the MEK/ERK pathway is not essential for the production of iNOS and IL-1β in macrophages (34). In the present study, MEC inhibited the LPS-induced phosphorylation of p38 and JNK, but not ERK, in a dose-dependent manner. However, the total MAPK levels were unchanged. Collectively, these results suggested that the activity of MAPKs, including p38 and JNK, rather than the expression of MAPKs, is the key regulatory mechanism underlying the MEC-mediated inhibition of inflammatory mediators.

NF-κB is the other major regulatory signaling molecule for inflammation. Following LPS stimulation, which leads to the phosphorylation and degradation of IκB in the cytosol, NF-κB subunits are freely translocated into the nucleus (35,36). The nuclear translocated NF-κB subunits, p65 and p50, regulate the production of various inflammatory mediators, including TNF-α, IL-6 and NO (37,38). In the present study, stimulation of macrophages by LPS led to the activation of NF-κB and MAPKs. However, MEC had no effect on the activity of NF-κB or on the degradation of IκB in the LPS-stimulated RAW 264.7 macrophages. NF-κB and MAPK signaling share TLR4 adaptor molecules and accessory molecules, including myeloid differentiation primary response 88 (MyD88), toll/IL-1 receptor domain-containing adapter-inducing interferon-β (TRIF), tumor-necrosis factor receptor-associated factor 6 (TRAF6)and IL receptor-associated kinase 1. However, the downstream signaling molecules of MyD88, TRIF and TRAF6 are quite different in the NF-κB and MAPK signaling pathways, therefore, the present study hypothesized that MEC may selectively regulate JNK and p38, but not the TLR4 accessory molecules.

The present study examined the regulation of inflammatory mediators by MEC in activated macrophages. Since macrophages are important in the pathogenesis of numerous inflammatory diseases, the MEC-mediated selective regulation of inflammatory mediators suggests that MEC may have therapeutic potential against inflammatory diseases. However, further studies are required to analyze the major components that are responsible for the reduction of inflammatory mediators and to elucidate the exact mechanism underlying the difference in the production of pro-inflammatory cytokines, IL-6 and TNF-α.

Acknowledgements

This study was supported by the National Research Foundation of Korea grant funded by the Korean government (Ministry of Education, Science and Technology; grant nos. NRF-2012R1A2A2A01047338 and NRF-2013R1A1A2062389).

References

1 

Fujiwara N and Kobayashi K: Macrophages in inflammation. Curr Drug Targets Inflamm Allergy. 4:281–286. 2005. View Article : Google Scholar

2 

Schroder K, Sweet MJ and Hume DA: Signal integration between IFNgamma and TLR signalling pathways in macrophages. Immunobiology. 211:511–524. 2006. View Article : Google Scholar : PubMed/NCBI

3 

Aderem A and Underhill DM: Mechanisms of phagocytosis in macrophages. Annu Rev Immunol. 17:593–623. 1999. View Article : Google Scholar

4 

Aderem A and Ulevitch RJ: Toll-like receptors in the induction of the innate immune response. Nature. 406:782–787. 2000. View Article : Google Scholar : PubMed/NCBI

5 

Cario E, Rosenberg IM, Brandwein SL, Beck PL, Reinecker HC and Podolsky DK: Lipopolysaccharide activates distinct signaling pathways in intestinal epithelial cell lines expressing Toll-like receptors. J Immunol. 164:966–972. 2000. View Article : Google Scholar : PubMed/NCBI

6 

Fujihara M, Muroi M, Tanamoto K, Suzuki T, Azuma H and Ikeda H: Molecular mechanisms of macrophage activation and deactivation by lipopolysaccharide: roles of the receptor complex. Pharmacol Ther. 100:171–194. 2003. View Article : Google Scholar : PubMed/NCBI

7 

Hanada T and Yoshimura A: Regulation of cytokine signaling and inflammation. Cytokine Growth Factor Rev. 13:413–421. 2002. View Article : Google Scholar : PubMed/NCBI

8 

Nathan C and Xie QW: Regulation of biosynthesis of nitric oxide. J Biol Chem. 269:13725–13728. 1994.PubMed/NCBI

9 

MacMicking J, Xie QW and Nathan C: Nitric oxide and macrophage function. Annu Rev Immunol. 15:323–350. 1997. View Article : Google Scholar : PubMed/NCBI

10 

Clancy RM, Amin AR and Abramson SB: The role of nitric oxide in inflammation and immunity. Arthritis Rheum. 41:1141–1151. 1998. View Article : Google Scholar : PubMed/NCBI

11 

Moncada S: Nitric oxide: discovery and impact on clinical medicine. J R Soc Med. 92:164–169. 1999.PubMed/NCBI

12 

Dinarello CA: Anti-inflammatory agents: present and future. Cell. 140:935–950. 2010. View Article : Google Scholar : PubMed/NCBI

13 

Eke T, Al-Husainy S and Raynor MK: The spectrum of ocular inflammation caused by Euphorbia plant sap. Arch Ophthalmol. 18:13–16. 2000. View Article : Google Scholar

14 

Lazarini CA, Uema AH, Brandão GM, Guimarães AP and Bernardi MM: Croton zehntneri essential oil: effects on behavioral models related to depression and anxiety. Phytomedicine. 7:477–481. 2000. View Article : Google Scholar : PubMed/NCBI

15 

Thyagarajan SP, Subramanian S, Thirunalasundari T, Venkateswaran PS and Blumberg BS: Effect of Phyllanthus amarus on chronic carriers of hepatitis B virus. Lancet. 1:764–766. 1988. View Article : Google Scholar : PubMed/NCBI

16 

Suarez AI, Compagnone RS, Salazar-Bookaman MM, et al: Antinociceptive and anti-inflammatory effects of Croton malambo bark aqueous extract. J Ethnopharmacol. 88:11–14. 2003. View Article : Google Scholar : PubMed/NCBI

17 

Pereira U, Garcia-Le Gal C, Le Gal G, et al: Effects of sangre de drago in an in vitro model of cutaneous neurogenic inflammation. Exp Dermatol. 19:796–799. 2010. View Article : Google Scholar : PubMed/NCBI

18 

Risco E, Ghia F, Vila R, Iglesias J, Alvarez E and Cañigueral S: Immunomodulatory activity and chemical characterisation of sangre de drago (dragon’s blood) from Croton lechleri. Planta Med. 69:785–794. 2003.PubMed/NCBI

19 

Nathan C and Xie QW: Nitric oxide synthases: roles, tolls, and controls. Cell. 78:915–918. 1994. View Article : Google Scholar : PubMed/NCBI

20 

Szabo C and Thiemermann C: Regulation of the expression of the inducible isoform of nitric oxide synthase. Adv Pharmacol. 34:113–153. 1995. View Article : Google Scholar : PubMed/NCBI

21 

Bosca L, Zeini M, Traves PG and Hortelano S: Nitric oxide and cell viability in inflammatory cells: a role for NO in macrophage function and fate. Toxicology. 208:249–258. 2005. View Article : Google Scholar : PubMed/NCBI

22 

Kröncke KD, Fehsel K and Kolb-Bachofen V: Inducible nitric oxide synthase in human diseases. Clin Exp Immunol. 113:147–156. 1998.

23 

Guadagni F, Ferroni P, Palmirotta R, Portarena I, Formica V and Roselli M: Review. TNF/VEGF cross-talk in chronic inflammation-related cancer initiation and progression: an early target in anticancer therapeutic strategy. In Vivo. 21:147–161. 2007.PubMed/NCBI

24 

Nishimoto N and Kishimoto T: Interleukin 6: from bench to bedside. Nat Clin Pract Rheumatol. 2:619–626. 2006. View Article : Google Scholar : PubMed/NCBI

25 

Kavanaugh A: Interleukin-6 inhibition and clinical efficacy in rheumatoid arthritis treatment - data from randomized clinical trials. Bull NYU Hosp Jt Dis. 65:S16–S20. 2007.PubMed/NCBI

26 

Straub RH, Harle P, Yamana S, et al: Anti-interleukin-6 receptor antibody therapy favors adrenal androgen secretion in patients with rheumatoid arthritis: a randomized, double-blind, placebo-controlled study. Arthritis Rheum. 54:1778–1785. 2006. View Article : Google Scholar

27 

Samavati L, Rastogi R, Du W, Huttemann M, Fite A and Franchi L: STAT3 tyrosine phosphorylation is critical for interleukin 1 beta and interleukin-6 production in response to lipopolysaccharide and live bacteria. Mol Immunol. 46:1867–1877. 2009. View Article : Google Scholar : PubMed/NCBI

28 

Zhu ZG, Jin H, Yu PJ, Tian YX, Zhang JJ and Wu SG: Mollugin inhibits the inflammatory response in lipopolysaccharide-stimulated RAW264.7 macrophages by blocking the janus kinase-signal transducers and activators of transcription signaling pathway. Biol Pharm Bull. 36:399–406. 2013. View Article : Google Scholar

29 

Lee C, Lim HK, Sakong J, Lee YS, Kim JR and Baek SH: Janus kinase-signal transducer and activator of transcription mediates phosphatidic acid-induced interleukin (IL)-1beta and IL-6 production. Mol Pharmacol. 69:1041–1047. 2006.PubMed/NCBI

30 

Stalinska K, Guzdek A, Rokicki M and Koj A: Transcription factors as targets of the anti-inflammatory treatment. A cell culture study with extracts from some Mediterranean diet plants. J Physiol Pharmaco. 56:S157–S169. 2005.

31 

Burk DR, Senechal-Willis P, Lopez LC, Hogue BG and Daskalova SM: Suppression of lipopolysaccharide-induced inflammatory responses in RAW 264.7 murine macrophages by aqueous extract of Clinopodium vulgare L. (Lamiaceae). J Ethnopharmacol. 126:397–405. 2009. View Article : Google Scholar : PubMed/NCBI

32 

Liu H, Bargouti M, Zughaier S, et al: Osteoinductive LIM mineralization protein-1 suppresses activation of NF-kappaB and selectively regulates MAPK pathways in pre-osteoclasts. Bone. 46:1328–1335. 2010. View Article : Google Scholar : PubMed/NCBI

33 

Shan J, Fu J, Zhao Z, et al: Chlorogenic acid inhibits lipopolysaccharide-induced cyclooxygenase-2 expression in RAW264.7 cells through suppressing NF-kappaB and JNK/AP-1 activation. Int Immunopharmacol. 9:1042–1048. 2009. View Article : Google Scholar : PubMed/NCBI

34 

Watters JJ, Sommer JA, Pfeiffer ZA, Prabhu U, Guerra AN and Bertics PJ: A differential role for the mitogen-activated protein kinases in lipopolysaccharide signaling: the MEK/ERK pathway is not essential for nitric oxide and interleukin 1beta production. J Biol Chem. 277:9077–9087. 2002. View Article : Google Scholar

35 

Karin M and Ben-Neriah Y: Phosphorylation meets ubiquitination: the control of NF-[kappa]B activity. Annu Rev Immunol. 18:621–663. 2000.PubMed/NCBI

36 

Karin M and Delhase M: The I kappa B kinase (IKK) and NF-kappa B: key elements of proinflammatory signalling. Semin Immunol. 12:85–98. 2000. View Article : Google Scholar : PubMed/NCBI

37 

Surh YJ, Chun KS, Cha HH, et al: Molecular mechanisms underlying chemopreventive activities of anti-inflammatory phytochemicals: down-regulation of COX-2 and iNOS through suppression of NF-kappa B activation. Mutat Res. 480–481:243–268. 2001.

38 

Verma IM, Stevenson JK, Schwarz EM, Van Antwerp D and Miyamoto S: Rel/NF-kappa B/I kappa B family: intimate tales of association and dissociation. Genes Dev. 9:2723–2735. 1995. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

November-2014
Volume 10 Issue 5

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Cho YC, Lee IS, Seo H, Ju A, Youn D, Kim Y, Choun J and Cho S: Methanol extracts of Euphorbia cooperi inhibit the production of inflammatory mediators by inhibiting the activation of c‑Jun N‑terminal kinase and p38 in murine macrophages. Mol Med Rep 10: 2663-2668, 2014
APA
Cho, Y., Lee, I., Seo, H., Ju, A., Youn, D., Kim, Y. ... Cho, S. (2014). Methanol extracts of Euphorbia cooperi inhibit the production of inflammatory mediators by inhibiting the activation of c‑Jun N‑terminal kinase and p38 in murine macrophages. Molecular Medicine Reports, 10, 2663-2668. https://doi.org/10.3892/mmr.2014.2560
MLA
Cho, Y., Lee, I., Seo, H., Ju, A., Youn, D., Kim, Y., Choun, J., Cho, S."Methanol extracts of Euphorbia cooperi inhibit the production of inflammatory mediators by inhibiting the activation of c‑Jun N‑terminal kinase and p38 in murine macrophages". Molecular Medicine Reports 10.5 (2014): 2663-2668.
Chicago
Cho, Y., Lee, I., Seo, H., Ju, A., Youn, D., Kim, Y., Choun, J., Cho, S."Methanol extracts of Euphorbia cooperi inhibit the production of inflammatory mediators by inhibiting the activation of c‑Jun N‑terminal kinase and p38 in murine macrophages". Molecular Medicine Reports 10, no. 5 (2014): 2663-2668. https://doi.org/10.3892/mmr.2014.2560