Anti-inflammatory effects of Calophyllum inophyllum L. in RAW264.7 cells

  • Authors:
    • Shih-Chang Tsai
    • Yu-Hsin Liang
    • Jo-Hua Chiang
    • Fon-Chang Liu
    • Wen-Hsin Lin
    • Shu-Jen Chang
    • Wei-Yu Lin
    • Chieh-Hsi Wu
    • Jing-Ru Weng
  • View Affiliations

  • Published online on: June 19, 2012     https://doi.org/10.3892/or.2012.1873
  • Pages: 1096-1102
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Calophyllum inophyllum L. has been used as folk medicine in the treatment of ocular burn and it has demonstrated potential to be an anti-inflammatory agent. The aim of this study was to explore the anti-inflammatory activities of an acetone extract of C. inophyllum L. leaves (CIL). The CIL extract was tested on lipopolysaccharide (LPS)-induced RAW 264.7 cells to evaluate the effect of CIL extract on the expression of nitric oxide (NO) and inducible nitric oxide synthase (iNOS). Results showed that the CIL extract markedly suppressed the LPS-induced production of nitric oxide, as well as the expression of iNOS, cyclooxygenase (COX)-2 and nuclear factor-kappaB (NF-κB) in a dose-dependent manner. LPS-induced microRNA (miR)-146a expression was inhibited by CIL extract, while miR-155 and miR-424 expression was not affected as demonstrated using quantitative RT-PCR analysis. Taken together, these observations show that CIL extract has anti-inflammatory effect, which extends the potential application for prevention of inflammatory diseases, and its mechanism may be partially associated with blocking COX-2 and iNOS of RAW 264.7 cells.

Introduction

Inflammation is a vital defense mechanism for the organism in response to pathogen stimuli. Monocytes and macrophages are dominant at the locations of lipopolysaccharide (LPS) induced inflammation. Macrophages become active upon LPS stimulation, and several cellular mediators such as tumor necrosis factor-α (TNF-α), IL-1β, nitric oxide (NO) and cyclooxygenase (COX)-2 are released to regulate inflammation (1).

NO is also known as a short-lived free radical and one of key cellular mediator for inflammatory responses. There are three isoforms of NO synthases (NOS) in tissues to generate NO (2). The neural NOS and endothelial NOS isoforms are constitutively expressed in select tissues. Inducible NOS (iNOS), a third member of the NOS family, is known to have beneficial effects in response to inflammatory stimuli. The expression level of iNOS is elevated in response to LPS via a variety of transcription factors, particularly nuclear factor-kappaB (NF-κB) (3). Cyclooxygenase (COX)-2, another cellular inflammatory mediator, is undetectable in most normal tissues (4). Upon LPS stimulation, COX-2 expression is induced rapidly and transiently at inflammatory sites (5). Furthermore, NF-κB plays vital roles in the coordination of the expression of pro-inflammatory mediators and cytokines, including iNOS (6) and COX-2 (7). Once cells are stimulated by LPS, NF-κB becomes activated, dissociates with IκB and then translocates from cytosol to nucleus, which leads to induction of NF-κB downstream genes via binding to the NF-κB response element (8). MicroRNAs function as non-coding RNA molecules that regulate the expression of target genes involved in a wide range of biological processes. It has been shown that microRNAs are found in almost all eukaryotic cells. Mature microRNAs bind to the 3′ untranslated regions of target genes and inhibit protein synthesis (9). Recently, microRNAs have been found to be involved in the growth and differentiation of immune cells (10). Previous studies showed that miR-146a and miR-155 were up-regulated in response to LPS stimulation. MiR-146 has been considered as a negative regulator in the innate immunity, while miR-155 regulated mature T cells and B cells (11). Moreover, miR-424 was found to be involved in the differentiation of macrophages (12).

Over-reactive and uncontrolled inflammatory responses can cause the tissues to remain in a chronic inflammatory status, which leads to a variety of diseases including rheumatoid arthritis, pulmonary fibrosis and even cancer (13). Previous studies have been shown that excessive NO production causes inflammation and carcinogenesis (14,15). Natural products have been used in drug discovery and development to regulate inflammation.

Calophyllum inophyllum Linn. (Guttiferae) is a medium to large tree distributed throughout Taiwan, India, and Australia (1619). The active constituents of C. inophyllum is well known for containing xanthone, flavone, and terpene derivatives, some of which exhibit antitumor (20,21), and anti-HIV activities (22,23). The oil of C. inophyllum enhanced the healing of Ocular burn (24). The present study was undertaken to evaluate the anti-inflammatory effects of C. inophyllum in macrophage cells under LPS exposure. Given that NO and COX-2 are specific to inflamed tissue, the inhibition of NO over-production and COX-2 expression is important for evaluating the effects of anti-inflammatory drugs. However, the effect of C. inophyllum L. leaf (CIL) extract on LPS-induced RAW264.7 cells is still unclear. The purpose of this study was to investigate the anti-inflammatory actions of CIL extract on NO production, COX-2 expression and translocation of NF-κB. In addition, we also provide evidence suggesting that CIL extract also suppressed LPS-induced miR-146a expression.

Materials and methods

Chemicals and reagents

Celebrex, acetone, dimethylsulfoxide (DMSO), lipopolysaccride (LPS), and N-(3-aminomethyl)-benzylacetamidine (1400W) were purchased from Sigma-Aldrich Corp. (St. Louis, MO, USA).

Preparation of the Calophyllum inophyllum Linn extracts

The leaves of C. inophyllum L. (Guttiferae) were collected in Ping Tung Hsieng, Taiwan, in October, 2008, and a voucher specimen (2008) has been deposited in the Department of Biological Science and Technology, China Medical University. The dried leaves of C. inophyllum (1 kg) was ground, extracted with acetone at room temperature, and concentrated under reduced pressure to afford a brown residue (90 g). The series concentrations of CIL extract were further diluted with DMSO.

Quantitative analysis of active compounds of C. inophyllum by HPLC

Before analysis by HPLC, CIL extract was filtered through a 0.2 μm Millipore filter, and then total volume of 20 μl was loaded into the HPLC column. External standards were prepared as concentration of 100 μg/ml in HPLC grade-methanol and used to calculate the concentration of examined compounds. Reverse phase HPLC was performed on a Perkin-Elmer HPLC system (Perkin-Elmer, Waltham, MA, USA) equipped with Perki-Elmer Series 200 pump, Perki-Elmer 785A UV/VIS detector and Perk-Elmer Series 200 autosampler. Separations were accomplished on LiChroCART 250-4 C18 HPLC-cartridge (5 μm; Merck, Whitehouse Station, NJ, USA). The separation conditions of HPLC analysis of examined compounds are described in Table I.

Table I

HPLC separation conditions for identifying marked components within CIL extract.a

Table I

HPLC separation conditions for identifying marked components within CIL extract.a

CompoundsMobile phaseWavelength (nm)RT (min)Contents (mg/g of CIL)
Amentoflavone MeOH:H3PO4 (0.11%, pH 2.2) = 25:7533010.3984.72±1.46
Oleanolic acid ACN:H3PO4 (0.11%, pH 2.2) = 28:7221510.151.05±0.01

a ACN, acetonitrile; RT, retention time. All samples were loaded at a total volume of 10 μl into the HPLC cartridge; a flow rate of 1.0 ml/min was used to perform HPLC analysis.

Cell culture and treatment

The murine macrophage RAW 264.7 cells were obtained from the Bioresource Collection and Research Center (BCRC, Hsinchu, Taiwan) and cultured in Dulbecco’s modified Eagle’s medium (Invitrogen, Carlsbad, CA, USA) supplemented with 10% heat-inactivated fetal bovine serum, 100 U/ml of penicillin, and 100 μg/ml of streptomycin. Cells were incubated with CIL extract as indicated for 1 h and then stimulated with 1 μg/ml lipopolysaccride (LPS) (Escherichia coli 011:B4, Sigma Chemical Co., St. Louis, MO, USA) for 24 h.

Cell viability assay

Cell viability of RAW 264.7 cells was measured by 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) as described elsewhere (25,26). Briefly, 1×104 RAW 264.7 cells/well were seeded in 96-well plates and incubated with various concentrations of CIL extract (0–20 μg/ml) at 37°C for 24 h and medium was completely removed. MTT was added to the cells, followed by incubation for 4 h at 37°C. After incubation, the medium was discarded and the formazan crystals in viable cells were dissolved in 100 μl of fresh DMSO for 10 min. The absorbance was calculated at 590 nm using a microplate autoreader (Molecular Devices, Sunnyvale, CA, USA). Relative cell viability was calculated by comparing the absorbance of the treated group to the LPS stimulated control group. All experiments were performed in triplicate.

Measurement of nitrite production

Nitric oxide (NO) production was determined by measurement of the accumulation of nitrite, the stable metabolite of NO in the culture medium. Nitrite was assayed colorimetrically after reaction with the Griess reagent as described previously (27). Briefly, 2×105 cells per well were seeded onto 96-well plates and then treated with CIL extract (0, 1, 2.5, 5, and 10 μg/ml) at 37°C for 1 h before stimulation with 1 μg/ml of LPS for 24 h in a final volume of 0.2 ml. The supernatant of LPS-induced RAW 264.7 cell cultures was mixed with an equal volume of Griess reagent (1% sulfanilamide and 0.1% naphthylenediamine in 1 N hydrochloric acid) in a 96-well plate. Nitrite concentrations were calculated by comparison with OD550 of standard solutions of sodium nitrite in culture medium. All determinations were performed in triplicate.

Quantification of mRNA and miRNA expression level by quantitative real-time PCR

Total RNA was extracted using TRIzol reagent (Invitrogen). Total RNA (1 μg) was heated at 70°C for 10 min and reversely transcribed using reverse transcriptase 200 U (Promega, Madison, WI, USA). The mixture was then incubated at 37°C for 60 min, heated at 95°C for 10 min and stored at −20°C until use. Real-time PCR was performed 40 cycles with primers for iNOS and β-actin as an internal control using the ABI PRISM 7300 Sequence Detector System (Applied Biosystems, Foster City, CA, USA). Cycles consisted of 30 sec of denaturation at 95°C, 30 sec of annealing at 60°C, 1 min of extension at 72°C, followed by 10 min of elongation at 72°C. Data were collected by the Sequence Detection Software (SDS; Version 1.3.1, Applied Biosystems) and analyzed using the threshold cycle relative quantification method. Primers were designed with computer assistance according to the gene bank. The sequence of the primers are as follows; iNOS sense primer 5′-CAT CAA CCA GTA TTA TGG CTC CT-3′, and iNOS anti-sense 5′-TCC TGT TGT TTC TAT TTC CTT TGT T-3′; β-actin sense 5′-CTA AGG CCA ACC GTG AAA AG-3′, and β-actin antisense 5′-ACC AGA GGC ATA CAG GGA CA-3′. The cycle threshold (Ct) values were determined in at least three independent experiments for each sample. Results were normalized to the endogenous gene β-actin.

MicroRNA quantification

TaqMan miRNA assays (Applied Biosystems) were used to quantify mature miRNA of miR-146a, miR-155 and miR-424. RNU6B was used as a reference gene control. Quantitative assays were performed in the 7900HT system according to the manufacturer’s instructions (Applied Biosystems). Briefly, 10 ng of RNA samples extracted either normal parts or tumor of same oral patient were mixed and reacted with reverse transcription master mix reagents (Applied Biosystems) at 16°C for 30 min, 42°C for 30 min, and 85°C for 5 min. Each of PCR cycle contained the denaturation step at 95°C for 15 sec and the extension step at 60°C for 1 min for a total of 40 cycles. miRNA expression values were normalized using RNU6B following the 2−ΔΔCt method and analyzed by using ABI 7900HT SDS 2.2 software. All reactions were run in triplicate.

Transient transfection

Transfection followed the manufacturer’s protocol (Thermo Scientific Open Biosystems, Huntsville, AL, USA). Briefly, 3×105 RAW 264.7 cells/well were incubated in 6-well plates overnight. Medium was changed into the DMEM without 10% heat-inactivated fetal bovine serum, 100 U/ml of penicillin, and 100 μg/ml of streptomycin before transfection. Solution A (2 μg of DNA and 50 μl of DMEM) was mixed with solution B (10 μl of Arrest-In transfection reagent and 50 μl of DMEM) at room temperature for 20 min. Transfection reagents were equally added into cells. Cells were harvested after 48-h transfection.

Luciferase reporter gene assays

pCOX-2-LUC plasmid was used to quantify COX-2 promoter activity. pRL-CMV, a renilla luciferase reporter plasmid under the control of the cytomegavirus promoter (Promega), was used as an internal control to normalize the reporter gene activity. Plasmids were transfected into RAW264.7 cells. After 24-h transfection, cells were treated with CIL extract for 1 h followed by stimulation with LPS and analyzed the luciferase activity after 48-h transfection. The luciferase activity was determined by a luminometer using a dual-luciferase reporter assay system (Promega) according to the instructions of the manufacturer. Briefly, cells were lysed with 1× PLB (passive lysis buffer) for 15 min. PLB lysate (20 μl) was added into a 96-well plate and mixed with 80 μl of LAR II (luciferase assay substrate in luciferase assay buffer II) to measure luciferase activity by using the SpectraMax L spectrometer (Molecular Devices, Sunnyvale, CA, USA). The values were normalized with the measurement of renilla luciferase activity.

Preparation of nuclear and cytosolic extract

The nuclear extract was prepared as described previously (28). Briefly, 5×105 RAW 264.7 cells were incubated with or without concentrations of CIL extract for 1 h, and then treated with LPS (100 ng/ml) for 30 min. After LPS treatment for 24 h, cells were harvested, washed with ice-cold PBS, and then centrifuged at 2500 g for 5 min at 4°C. Cell pellets were added to 100 μl lysis buffer (10 mM HEPES pH 7.9, 10 mM KCl, 0.1 mM EDTA, 0.5% Nonidet-P 40, 1 mM dithiothreitol, 0.5 mM PMSF) and vortexed mildly. Samples were incubated for 10 min on ice and centrifuged at 2500 g for 5 min at 4°C. The supernatant was collected as a cytosolic fraction. Pellets containing crude nuclei were resuspended in 100 μl extraction buffer (20 mM HEPES pH 7.9, 400 mM NaCl, 1 mM EDTA, 1 mM dithiothreitol, 1 mM PMSF) and incubated for 30 min on ice, and centrifuged at 15,000 g for 10 min. The supernatant containing nuclear extracts was collected and stored at −80°C until required.

Western blot analysis

Proteins were separated by SDS-PAGE and transferred onto PVDF (Millipore, Billerica, MA, USA) as described previously (2931). Nonspecific binding on the nitrocellulose filter paper was minimized with a blocking buffer containing 5% non-fat dry milk and 0.1% Tween-20 in PBS. The membrane was incubated with specific primary antibodies to COX-2 or NF-κB (Abcam, Cambridge, UK) followed by incubation with horseradish peroxidase-conjugated goat anti-rabbit antibody (1:7000 dilution, Abcam). For internal controls, the same membranes were incubated with mouse anti-β-actin and anti-PCNA for 1 h followed by incubation with horseradish-peroxidase-linked goat anti-mouse IgG for 1 h. Reactive bands were visualized with an enhanced chemiluminescence system (Amersham Biosciences, Arlington Heights, IL). The intensity of the bands was scanned and quantified with Adobe Photoshop software.

Statistical analysis

Data are expressed as mean ± standard deviations (SD) from three different experiments. Statistical analysis was carried out using the Student’s t-test. It was considered statistically significant at *p<0.05, and **p<0.01.

Results and Discussion

Effect of CIL extract on cell proliferation of LPS-induced RAW264.7 cells

After 24-h treatment with CIL extract at the indicated concentrations (0–25 μg/ml) in RAW264.7 cells, CIL extract was found to inhibit RAW264.7 cell proliferation in a dose-dependent manner, and the IC50 value of CIL extract for 24 h was 14 μg/ml (Fig. 1). Therefore, sample treatments between 1 and 10 μg/ml were used in the subsequent experiments.

CIL extract inhibits LPS-induced NO production and iNOS gene expression in RAW 264.7 macrophage cells

We investigated the inhibitory effect of CIL extract on LPS-induced NO production in RAW264.7 cells. Cells were preincubated with CIL extract at the indicated concentrations for 1 h and then stimulated with LPS for 23 h. Supernatants were collected for determination of nitrite production. CIL extract inhibited NO production in RAW264.7 cells in a dose-dependent manner as compared to controls, and the anti-inflammatory agent N-(3-aminomethyl)-benzylacetamidine (1400W) as the positive control (Fig. 2A). Quantitative RT-PCR showed that the iNOS mRNA expression was almost undetectable in unstimulated cells but were markedly augmented by LPS. Upon LPS stimulation, iNOS expression was induced ~25-fold, and CIL extract blocked iNOS expression which was induced by LPS 7.6-fold in RAW264.7 cells (Fig. 2B). The LPS-induced iNOS expression was inhibited 11.5-fold by the iNOS inhibitor 1400W at 10 μM (Fig. 2B).

CIL extract suppresses mRNA levels of COX-2

QPCR was performed to determine whether the inhibitory effects of the CIL extract on the pro-inflammatory mediators (NO) were related to the modulation of the expression of iNOS and COX-2. After the transient transfection of RAW 264.7 cells with pCOX-2-LUC and pRL-CMV, the expressions of firefly luciferase and Renilla luciferase, respectively, were used to quantify the COX-2 promoter activity. RAW264.7 cells were treated with the various concentrations of CIL extract (0, 1, 2.5, 5 and 10 μg/ml). As shown in Fig. 1, CIL extract did not exhibit cytotoxicity against RAW264.7 cells <2.5 μg/ml. The basal level of COX-2 promoter activity was set in the absence of LPS stimulation. LPS significantly induced up to 203% in the production of COX-2. CIL extract significantly reduced to 128% at 1 μg/ml, and 82% at 2.5 μg/ml in the production of LPS-induced COX-2 promoter activity (Fig. 3).

CIL extract inhibits LPS-induced NF-κB activation

Since p65 and p50 are the major components of NF-κB activated by LPS in the macrophage (32,33), the levels of NF-κB/p65 in the nuclear extract were determined by western blot analysis (Fig. 4A). RAW 264.7 cells were incubated with LPS in the presence or absence of the CIL extract with various concentrations for 24 h. The expression of NF-κB/p65 in the nucleus was markedly increased upon exposure to LPS alone 2.3-fold, but the extract or CIL extract inhibited LPS-mediated nuclear translocation of NF-κB/p65 in a dose-dependent manner (Fig. 4B) (p<0.05), indicating that CIL extract could inhibit the nuclear translocation of NF-κB/p65.

CIL extract attenuates LPS-induced microRNA-146a, 155, and 424 expressions

When RAW264.7 cells were exposed to LPS, expressions of microRNA-146a and miR-155 were up-regulated 6.97- and 77.23-fold, respectively (Fig. 5A and B). After treatment with 10 μg/ml CIL extract for 24 h, the expression of microRNA-146a was reduced to 4.26-fold (Fig. 5A). However, there was no significant change for LPS or CIL extract in the miR-424 expression level (Fig. 5C).

Quality of extraction procedure of CIL extract by HPLC analysis

Two of the marked components of CIL extract, including amentoflavone and oleanolic acid, were identified by HPLC analysis to be indicator compounds for quality check of extraction procedure of each batch (Fig. 6). Using HPLC, the contents of amentoflavone and oleanolic acid were calculated to be 84.72 and 1.05 mg/g of CIL extract, respectively (Table I).

Macrophage activation is important to the progression of multiple diseases through the release of inflammatory mediators. Lipopolysaccharide (LPS)-induced RAW264.7 macrophages are widely used in vitro, because LPS is a pathogen that triggers toll-like receptor 4 (TLR4) and activates various inflammatory signals (34). Natural products have played a significant role in drug discovery and development, especially agents against several diseases that have existed from antiquity to the present. Since inflammation is closely linked to the promotion of certain tumors, substances with potent anti-inflammatory activities are anticipated to exert chemopreventive effects on carcinogenesis (35).

In the present study, we prepared acetone extract from CIL extract and examined its effects on the LPS-induced inflammation in a murine macrophage cell line RAW 264.7 model. First, the cytotoxicity of CIL extract in RAW 264.7 cells were evaluated by MTT assay, and it was observed that CIL extract did not affect cell viability <2.5 μg/ml. Many lines of evidence have indicated that NO is a potent proinflammatory mediator and may have a multi-faceted role in mutagenesis and carcinogenesis (35). The massive amounts of NO produced in response to bacterial LPS or cytokines play an important role in the inflammatory condition. Improper activation or upregulation of iNOS or COX-2 has been shown to be associated with the pathophysiologies of certain types of human cancer as well as inflammatory disorders. Therefore, aberrant or excessive expression of iNOS is often implicated in the oncogenesis and pathogenesis of cancer. Indeed, we found that CIL extract demonstrated the inhibition of the NO production in LPS-stimulated RAW 264.7 macrophages. The mRNA expression levels of iNOS in LPS-stimulated cells were examined by western blot analyses and quantitative RT-PCR compared with the specific iNOS inhibitor 1400W. Our results suggest that CIL extract blocks the transcription level of iNOS. Moreover, CIL extract at 1 μg/ml also suppressed the LPS-induced COX2 promoter activity in RAW 264.7 cells. Taken together, CIL extract suppressed the production of NO and downregulated the expression of iNOS and COX-2 in LPS-induced RAW 264.7 cells.

NF-κB is a transcription factor that plays a critical role in inflammatory and immune responses. It is present in the cytoplasm, binding to the inhibitory protein IκB in unstimulated cells (36,37). When the cells are exposed to the stimulants such as LPS, IκB is phosphorylated and liberates NF-κB, resulting in NF-κB translocation into the nucleus. Nuclear NF-κB then binds to the promoters of pro-inflammatory mediators, resulting in the induction of their gene expression (38). Here, we have elucidated that CIL extract diminished the LPS-induced NF-κB nuclear translocation in RAW 264.7 cells by western blot analysis. The nuclear translocation and DNA binding of NF-κB is essential for the LPS-mediated NO production and COX2 expression (39). These findings suggest that CIL extract may prevent inflammation by suppressing the NF-κB-mediated inflammatory gene.

Up to date, miRNAs have been demonstrated to be dysregulated in cancer (40) and aberrantly expressed in such inflammatory diseases as rheumatoid arthritis (41). Thus, miRNAs may form a key link between inflammation and cancer; however, the induction of specific miRNAs, including miR-146a and miR-155, as a key step in tumor progression is still unclear. While miR-146a has previously been reported in response to various microbial components and cytokines, much is still not known about its biological significance. In this study, miR-146a expression was induced by LPS stimulation in RAW 264.7 cells. Consistent with previous findings, exposure of THP-1 monocytes to various bacterial inflammatory insults, such as LPS and endotoxin, resulted in rapid and continuous expression of mature miR-146a (42). Our results show that LPS-induced miR-146a expression was inhibited by the CIL extract. Notably, during LPS stimulation, several miRNAs, including miR-146a, miR-155 and miR-424 are upregulated. The biological significance of miR-155 and miR-424 to LPS-induced inflammation has been widely investigated. The new data are focused on the expression of miR-146a, miR-155 and miR-424 in LPS-induced inflammatory responses. CIL extract reduced miR-146a expression, instead of miR-155 and miR-424 expression. This finding highlights the importance of further studies on whether miR-146a and miR-155 can be used as a therapeutic intervention for controlling immune response and defining the role of miR-146a, miR-155 and miR-424 in LPS-induced inflammatory responses.

Furthermore, amentoflavone and oleanolic acid are two major components isolated from CIL extract (20,40,43). Flavonoids are naturally occurring polyphenolic compounds that have many biological properties, including antioxidative, anti-inflammatory and neuroprotective effects (44). Oleanolic acid is a triterpenoid compound that is widely found in vegetables, medicinal herbs, and other plants. It has been shown that oleanolic acid has potent antioxidant and anti-inflammatory effects (45,46).

Collectively, we have demonstrated that CIL extract inhibits LPS-induced NO production and iNOS expression which mediated through the inhibition of NF-κB activation in RAW 264.7 macrophages. CIL extract exerts its potent anti-inflammatory activity by suppressing COX-2 promoter activity and expression. Our results demonstrate the strong anti-inflammatory properties of CIL extract by inhibition of iNOS and COX-2 expression as well as miR-146a expression.

Acknowledgements

This study was supported by grants from the China Medical University and National Science Council (NSC98-2815-C-039-080-B) (CMU96-116).

Abbreviations:

CIL

Calophyllum inophyllum L. leaves

NF-κB

nuclear factor kappaB

NO

nitric oxide

DMSO

dimetylsulfoxide

miRNA

microRNA

1400W

N-(3-aminomethyl)-benzylacetamidine

COX-2

cyclooxygenase-2

LPS

lipopolysaccride

References

1 

Guha M and Mackman N: LPS induction of gene expression in human monocytes. Cell Signal. 13:85–94. 2001. View Article : Google Scholar : PubMed/NCBI

2 

Kleinert H, Schwarz PM and Forstermann U: Regulation of the expression of inducible nitric oxide synthase. Biol Chem. 384:1343–1364. 2003. View Article : Google Scholar : PubMed/NCBI

3 

Goldring CE, Reveneau S, Pinard D and Jeannin JF: Hyporesponsiveness to lipopolysaccharide alters the composition of NF-kappaB binding to the regulatory regions of inducible nitric oxide synthase gene. Eur J Immunol. 28:2960–2970. 1998. View Article : Google Scholar : PubMed/NCBI

4 

Turini ME and DuBois RN: Cyclooxygenase-2: a therapeutic target. Annu Rev Med. 53:35–57. 2002. View Article : Google Scholar : PubMed/NCBI

5 

Williams JA and Shacter E: Regulation of macrophage cytokine production by prostaglandin E2. Distinct roles of cyclooxygenase-1 and -2. J Biol Chem. 272:25693–25699. 1997. View Article : Google Scholar : PubMed/NCBI

6 

Wang H, Gao X, Fukumoto S, Tademoto S, Sato K and Hirai K: Post-isolation inducible nitric oxide synthase gene expression due to collagenase buffer perfusion and characterization of the gene regulation in primary cultured murine hepatocytes. J Biochem. 124:892–899. 1998. View Article : Google Scholar

7 

Baeuerle PA and Baltimore D: NF-kappa B: ten years after. Cell. 87:13–20. 1996.PubMed/NCBI

8 

Rothwarf DM and Karin M: The NF-kappa B activation pathway: a paradigm in information transfer from membrane to nucleus. Sci STKE. 1999:RE11999.PubMed/NCBI

9 

Bartel DP: MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 116:281–297. 2004. View Article : Google Scholar : PubMed/NCBI

10 

Hobert O: Gene regulation by transcription factors and microRNAs. Science. 319:1785–1786. 2008. View Article : Google Scholar : PubMed/NCBI

11 

Pedersen I and David M: MicroRNAs in the immune response. Cytokine. 43:391–394. 2008. View Article : Google Scholar : PubMed/NCBI

12 

Rosa A, Ballarino M, Sorrentino A, et al: The interplay between the master transcription factor PU. 1 and miR-424 regulates human monocyte/macrophage differentiation. Proc Natl Acad Sci USA. 104:19849–19854. 2007. View Article : Google Scholar : PubMed/NCBI

13 

Doggrell SA: Inflammation, the key to much pathology. Drug News Perspect. 18:531–539. 2005.PubMed/NCBI

14 

Moon TC, Kim JC, Song SE, et al: Saucerneol D, a naturally occurring sesquilignan, inhibits LPS-induced iNOS expression in RAW264.7 cells by blocking NF-kappaB and MAPK activation. Int Immunopharmacol. 8:1395–1400. 2008. View Article : Google Scholar : PubMed/NCBI

15 

Vodovotz Y, Lucia MS, Flanders KC, et al: Inducible nitric oxide synthase in tangle-bearing neurons of patients with Alzheimer’s disease. J Exp Med. 184:1425–1433. 1996.PubMed/NCBI

16 

Venkanna BK and Venkataramana Reddy C: Biodiesel production and optimization from Calophyllum inophyllum linn oil (honne oil) - a three stage method. Bioresour Technol. 100:5122–5125. 2009. View Article : Google Scholar : PubMed/NCBI

17 

Kawazu K, Ohigashi H and Mitsui T: The piscicidal constituents of Calophyllum inophyllum Linn. Tetrahedron Lett. 19:2383–2385. 1968. View Article : Google Scholar : PubMed/NCBI

18 

Govindachari TR, Viswanathan N, Pai BR, Rao R and Srinivasan M: Triterpenes of Calophyllum inophyllum Linn. Tetrahedron. 23:1901–1910. 1967. View Article : Google Scholar : PubMed/NCBI

19 

Arora RB, Mathur CN and Seth SD: Calophyllolide, a complex coumarin anticoagulant from Calophyllum inophyllum Linn. J Pharm Pharmacol. 14:534–535. 1962. View Article : Google Scholar : PubMed/NCBI

20 

Li YZ, Li ZL, Yin SL, et al: Triterpenoids from Calophyllum inophyllum and their growth inhibitory effects on human leukemia HL-60 cells. Fitoterapia. 81:586–589. 2010. View Article : Google Scholar : PubMed/NCBI

21 

Dai HF, Zeng YB, Xiao Q, Han Z, Zhao YX and Mei WL: Caloxanthones O and P: two new prenylated xanthones from Calophyllum inophyllum. Molecules. 15:606–612. 2010. View Article : Google Scholar : PubMed/NCBI

22 

Spino C, Dodier M and Sotheeswaran S: Anti-HIV coumarins from Calophyllum seed oil. Bioorg Med Chem Lett. 8:3475–3478. 1998. View Article : Google Scholar : PubMed/NCBI

23 

De Clercq E: Current lead natural products for the chemotherapy of human immunodeficiency virus (HIV) infection. Med Res Rev. 20:323–349. 2000.PubMed/NCBI

24 

Said T, Dutot M, Labbe A, Warnet JM and Rat P: Ocular burn: rinsing and healing with ionic marine solutions and vegetable oils. Ophthalmologica. 223:52–59. 2009. View Article : Google Scholar : PubMed/NCBI

25 

Yang JS, Hour MJ, Huang WW, Lin KL, Kuo SC and Chung JG: MJ-29 inhibits tubulin polymerization, induces mitotic arrest, and triggers apoptosis via cyclin-dependent kinase 1-mediated Bcl-2 phosphorylation in human leukemia U937 cells. J Pharmacol Exp Ther. 334:477–488. 2010. View Article : Google Scholar

26 

Wu PP, Liu KC, Huang WW, et al: Triptolide induces apoptosis in human adrenal cancer NCI-H295 cells through a mitochondrial-dependent pathway. Oncol Rep. 25:551–557. 2011.PubMed/NCBI

27 

Lin MW, Tsao LT, Huang LJ, et al: Inhibition of lipopolysaccharide-stimulated NO production by crotafuran B in RAW 264.7 macrophages involves the blockade of NF-kappaB activation through the increase in IkappaBalpha synthesis. Toxicol Appl Pharmacol. 210:108–115. 2006. View Article : Google Scholar : PubMed/NCBI

28 

Andrews NC and Faller DV: A rapid micropreparation technique for extraction of DNA-binding proteins from limiting numbers of mammalian cells. Nucleic Acids Res. 19:24991991. View Article : Google Scholar : PubMed/NCBI

29 

Tsai SC and Seto E: Regulation of histone deacetylase 2 by protein kinase CK2. J Biol Chem. 277:31826–31833. 2002. View Article : Google Scholar : PubMed/NCBI

30 

Chiang JH, Yang JS, Ma CY, et al: Danthron, an anthraquinone derivative, induces DNA damage and caspase cascades-mediated apoptosis in SNU-1 human gastric cancer cells through mitochondrial permeability transition pores and Bax-triggered pathways. Chem Res Toxicol. 24:20–29. 2011. View Article : Google Scholar

31 

Lu CC, Yang JS, Huang AC, et al: Chrysophanol induces necrosis through the production of ROS and alteration of ATP levels in J5 human liver cancer cells. Mol Nutr Food Res. 54:967–976. 2010. View Article : Google Scholar : PubMed/NCBI

32 

Jeon YJ, Han SB, Ahn KS and Kim HM: Differential activation of murine macrophages by angelan and LPS. Immunopharmacology. 49:275–284. 2000. View Article : Google Scholar : PubMed/NCBI

33 

Kuo CT, Chiang LL, Lee CN, et al: Induction of nitric oxide synthase in RAW 264.7 macrophages by lipoteichoic acid from Staphylococcus aureus: involvement of protein kinase C- and nuclear factor-κB-dependent mechanisms. J Biomed Sci. 10:136–145. 2003.PubMed/NCBI

34 

Denlinger LC, Fisette PL, Garis KA, et al: Regulation of inducible nitric oxide synthase expression by macrophage purinoreceptors and calcium. J Biol Chem. 271:337–342. 1996. View Article : Google Scholar : PubMed/NCBI

35 

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.

36 

Baldwin AS Jr: The NF-kappa B and I kappa B proteins: new discoveries and insights. Annu Rev Immunol. 14:649–683. 1996. View Article : Google Scholar : PubMed/NCBI

37 

Simeonidis S, Liang S, Chen G and Thanos D: Cloning and functional characterization of mouse IkappaBepsilon. Proc Natl Acad Sci USA. 94:14372–14377. 1997. View Article : Google Scholar

38 

Gilroy DW, Lawrence T, Perretti M and Rossi AG: Inflammatory resolution: new opportunities for drug discovery. Nat Rev Drug Discov. 3:401–416. 2004. View Article : Google Scholar : PubMed/NCBI

39 

Ghosh S and Karin M: Missing pieces in the NF-kappaB puzzle. Cell. 109(Suppl): S81–S96. 2002. View Article : Google Scholar : PubMed/NCBI

40 

Li Y, Vandenboom TG II, Wang Z, et al: miR-146a suppresses invasion of pancreatic cancer cells. Cancer Res. 70:1486–1495. 2010. View Article : Google Scholar : PubMed/NCBI

41 

Chan EK, Satoh M and Pauley KM: Contrast in aberrant microRNA expression in systemic lupus erythematosus and rheumatoid arthritis: is microRNA-146 all we need? Arthritis Rheum. 60:912–915. 2009. View Article : Google Scholar : PubMed/NCBI

42 

Nahid MA, Pauley KM, Satoh M and Chan EK: miR-146a is critical for endotoxin-induced tolerance: implication in innate immunity. J Biol Chem. 284:34590–34599. 2009. View Article : Google Scholar : PubMed/NCBI

43 

Li YZ, Li ZL, Hua HM, Li ZG and Liu MS: Studies on flavonoids from stems and leaves of Calophyllum inophyllum. Zhongguo Zhong Yao Za Zhi. 32:692–694. 2007.PubMed/NCBI

44 

Shin DH, Bae YC, Kim-Han JS, et al: Polyphenol amentoflavone affords neuroprotection against neonatal hypoxic-ischemic brain damage via multiple mechanisms. J Neurochem. 96:561–572. 2006. View Article : Google Scholar

45 

Pereira DA, Dalmarco JB, Wisniewski A Jr, Simionatto EL, Pizzolatti MG and Frode TS: Lotus corniculatus regulates the inflammation induced by bradykinin in a murine model of pleurisy. J Agric Food Chem. 59:2291–2298. 2011. View Article : Google Scholar : PubMed/NCBI

46 

Sung HY, Kang SW, Kim JL, et al: Oleanolic acid reduces markers of differentiation in 3T3-L1 adipocytes. Nutr Res. 30:831–839. 2010. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

September 2012
Volume 28 Issue 3

Print ISSN: 1021-335X
Online ISSN:1791-2431

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Tsai S, Liang Y, Chiang J, Liu F, Lin W, Chang S, Lin W, Wu C and Weng J: Anti-inflammatory effects of Calophyllum inophyllum L. in RAW264.7 cells. Oncol Rep 28: 1096-1102, 2012
APA
Tsai, S., Liang, Y., Chiang, J., Liu, F., Lin, W., Chang, S. ... Weng, J. (2012). Anti-inflammatory effects of Calophyllum inophyllum L. in RAW264.7 cells. Oncology Reports, 28, 1096-1102. https://doi.org/10.3892/or.2012.1873
MLA
Tsai, S., Liang, Y., Chiang, J., Liu, F., Lin, W., Chang, S., Lin, W., Wu, C., Weng, J."Anti-inflammatory effects of Calophyllum inophyllum L. in RAW264.7 cells". Oncology Reports 28.3 (2012): 1096-1102.
Chicago
Tsai, S., Liang, Y., Chiang, J., Liu, F., Lin, W., Chang, S., Lin, W., Wu, C., Weng, J."Anti-inflammatory effects of Calophyllum inophyllum L. in RAW264.7 cells". Oncology Reports 28, no. 3 (2012): 1096-1102. https://doi.org/10.3892/or.2012.1873