Basic Research Open Access
Copyright ©The Author(s) 2005. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Gastroenterol. Oct 7, 2005; 11(37): 5782-5786
Published online Oct 7, 2005. doi: 10.3748/wjg.v11.i37.5782
Effect of Astragalus complanatus flavonoid on anti-liver fibrosis in rats
Chun-Yu Liu, Department of Pharmacology, Medical College of Suzhou University, Suzhou 215007, Jiangsu Province, China
Zhen-Lun Gu, Suzhou Institute of Chinese Meteria Medica, Suzhou University, Suzhou 215007, Jiangsu Province, China
Wen-Xuan Zhou, Ci-Yi Guo, The Hong Kong Association for Health Care Ltd, Hong Kong, China
Author contributions: All authors contributed equally to the work.
Correspondence to: Professor Zhen-Lun Gu, Suzhou Institute of Chinese Meteria Medica, Suzhou University, Suzhou 215007, Jiangsu Province, China. zhenlungu.2003@163.com
Telephone: +86-512-65190599 Fax: +86-512-65190599
Received: August 16, 2004
Revised: October 2, 2004
Accepted: October 7, 2004
Published online: October 7, 2005

Abstract

AIM: To observe the anti-liver fibrosis effect of Astragalus complanatus flavonoids (ACF) in rats.

METHODS: The liver fibrosis model in rats was established by injecting interperitoneally 0.2 mL/100 g 0.5% dimethylnitrosamine, thrice a week. Meanwhile, the rats were administered ACF (30, 60, 120 mg/kg) or colchicine (0.1 mg/kg) once a day for 1 mo. Serum N-propeptide of type I procollagen (PINP) and type III procollagen (PIIINP) was measured using ELISA. Malondialdehyde (MDA) and superoxide dismutase (SOD) in hepatic tissue were evaluated. Matrix metal protease-1 (MMP-1) mRNA expression was assayed by RT-PCR and the protein expression of tissue inhibitor of metal protease-1 (TIMP-1) was analyzed by immunohistochemistry.

RESULTS: In the ACF groups, SOD activity increased and MDA content decreased in comparison to the liver fibrosis model group. The serum PINP and PIIINP contents in ACF-2 and -3 group decreased compared to those in model group. In ACF-2 and -3 group, the expression of MMP-1 mRNA increased significantly and the protein expression of TIMP-1 decreased compared to that in model group.

CONCLUSION: The antifibrotic mechanisms of ACF are associated with its influence on lipid peroxidation and collagen synthesis and degradation.

Key Words: Astragalus complanatus, Liver fibrosis, N-propeptide of type I procollagen, N-propeptide of type III procollagen, Malondialdehyde, Superoxide dismutase, Matrix metal protease-1, Tissue inhibitor of metal protease-1



INTRODUCTON

Fibrosis is a common manifestation in most chronic liver diseases. The progression of hepatic fibrosis often leads to cirrhosis and is associated with liver cancer. Therefore, interrupting or reversing hepatic fibrosis is very important in preventing cirrhosis. Semen Astragali (SA) is a tonic traditional Chinese medicine for liver and kidney. We previously reported that SA has a protective effect on liver[1]. Our further investigation has revealed that the flavonoids extracted from SA are useful in preventing rat liver fibrosis induced by dimethylnitrosamine (DMN)[2]. However, how Astragalus complanatus flavonoids (ACF) interrupt the formation of liver fibrosis remains unknown. This study was designed to observe the anti-liver fibrosis effect and mechanisms of DMN in rats.

MATERIALS AND METHODS
Materials

Astragalus complanatus (AC) was obtained from Leiyunshang Pharmaceutical Corporation (Suzhou, China) and identified by a chief pharmacist (Suzhou Drug Administration Bureau). Colchicine was obtained from Kunming Pharmaceutical Corporation (Kunming, China). DMN was purchased from Shanghai Chemical Company (Shanghai, China). All other chemicals were of analytical grade.

Preparation of ACF

After being extracted with ligarine, the seed powder of AC was extracted with 80% ethyl acetate (EtOH). The 80% EtOH extract was further subjected to ZTC-5 adsorptive resin (Nankai University) column and eluented with water, 50% and 95% EtOH to afford three fractions. The 50% EtOH fraction was dried under vacuum, recrystallized with EtOH and ACF was obtained as a yellow powder containing complanatuside, neocomplanatuside, myricomplanoside, rhomnocitrin-3-O-b-D-glucoside and astragalin, etc. The complanatuside content in ACF was about 20% by RP-HPLC and determined on a Shim-pack VP-ODS analytical column (4.6 mm150 mm, 5 mm) with the mobile phase consisting acetonitrile-1% acetic acid (18:82) as eluent at the flow rate of 1 mL/min. The UV detection was set at 266 nm. The total ACF was 80%, determined at 266 nm by UV spectroscopy with complanatuside as a standard substance.

Animals

Sixty male Wistar rats weighing 130-160 g were housed in conventional cages with free access to water and rodent chow at 20-22 °C with a 12-h light-dark cycle. All procedures involving the use of laboratory animals were in accordance with National Institutes of Health Guidelines. Rat liver fibrosis was produced according to the established protocol[3]. Except for the normal group, rats in other groups were injected interperitoneally 5 g/L DMN (0.2 mL/100 g, thrice a week). Meanwhile, the rats were administered 30 mg/kg ACF-1, 60 mg/kg ACF-2, 120 mg/kg ACF-3 or 0.1 mg/kg colchicine, respectively, once a day for 1 mo. The doses of ACF and its course of treatment were based on a pilot experiment. The rats in the normal control group and DMN model group received saline by gavage instead of ACF. After 4 wk of ACF treatment, all rats were anesthetized, sera were collected and liver tissues were removed rapidly. Liver tissue samples were immediately rinsed in sterile saline, snap-frozen in liquid nitrogen, and stored at -70 °C until analysis.

MDA and SOD determination

The livers were homogenized at 100 g/L in cold saline at 0-4 °C using a glass-potter-type homogenizer at 500-800 r/min in ice. The homogenates were centrifuged at 100 000 g for 20 min at 4 °C to obtain supernatants of tissue homogenates. The resultant supernatants were immediately pipetted for measuring the content of malondialdehyde (MDA) and the enzymatic activity of superoxide dismutase (SOD) by the test kits. Data were expressed as nanomolar MDA per gram wet liver tissue and units SOD per milligram wet liver tissue.

PINP and PIIINP determination

Serum N-propeptide of type I procollagen (PINP) and N-propeptide of type III procollagen (PIIINP) were measured by ELISA (LIFEKEY BioMeditech Corporation).

Immunohistochemistry

Five-micrometer-thick sections of formalin-fixed and paraffin-embedded livers were processed routinely. Tissue inhibitor of metal protease-1 (TIMP-1) was stained with polyclonal antibody (Santa Cruz Biotechnology, Inc.). Briefly, sections were deparaffinized in xylene and hydrated in a series of graded alcohol, and endogenous peroxidase was blocked with 5% hydrogen peroxide. The sections were then incubated with primary antibodies (1:50) against TIMP-1 at 37 °C for 1 h, followed by incubation with goat anti-rabbit IgG conjugated-horseradish peroxidase (1:200). The signals were visualized by ABC immunostaining system. The images were captured and digitized with an image analysis system (Jieda 501, version 3.10). The TIMP-1 positive cell ratios were automatically evaluated. For each specimen at least five fields were analyzed at 20 magnification.

RT-PCR reaction

Semiquantitative RT-PCR with b-actin as internal control was performed to examine the expression of matrix metal protease-1 (MMP-1) mRNA. Total RNA was extracted from liver by TRIzol method (GIBCO specification). The first-strand cDNA was synthesized from the total RNA using 40 mL reverse transcription mixture containing 2 mg total RNA, 1.0 mg M-MLV reverse transcriptase (Promega) and random primers (Sangon, Shanghai), 10 mmol/L dNTP. Reverse transcription reaction was preformed at 37 °C for 1 h. At the end of reverse transcription, the mixture was heated at 95 °C for 5 min and immediately cooled in ice for 5 min.

PCR amplification was carried out with 2 mL cDNA product in a 40 mL reaction volume containing 10 pmoL of each specific oligonucleotide primer (Table 1), 10 mmol/L dNTP, and 2.5 IU Taq DNA polymerase. Amplification was performed for 30 cycles of denaturation (at 94 °C for 30 s), annealing (at 58 °C for 30 s), and elongation (at 72 °C for 50 s). PCR products were electrophoresed on 20 g/L agarose gel and visualized by ethidium bromide staining under UV light. This semiquantitative measure was expressed as ratios compared to b-actin.

Table 1 Primers for PCR amplification of MMP-1 and b-actin mRNA.
PrimernamePrimer sequenceProductSize/bpGenBankacc. no.
MMP-1Sense: 5’-TACTAGCCACCGCTTCTTCC-3’ (192–211)311U53605
Anti-sense: 5’-GGCCAGAAATAGCTGAATGC-3’ (483–502)
β-actinSense: 5’-CTGTGCCCATCTATGAGGGT-3’ (485–504)199NM-031144
Anti-sense: 5’-GCCATCTCTTGCTCGAAGTC-3’ (664–683)
Statistical analysis

All numerical data were expressed as mean±D. Statistical analyses were done using the SPSS for Window release 11.0 (SPSS Inc., USA). The parameters were evaluated with one-way ANOVA and Dunnett’s t test. P < 0.05 was considered statistically significant.

RESULTS
Effect of ACF on MDA and SOD in liver tissue of rats

As shown in Figure 1A, the SOD activity decreased and the MDA content increased significantly in model group in comparison to those in normal group (P < 0.01, P < 0.01). However, the SOD activity in ACF groups increased and MDA contents decreased in comparison to those in model group (P < 0.01).

Figure 1
Figure 1 Effects of ACF on SOD activity and MDA content in rat liver tissue (A) and on PINP and PIIINP contents in rats serum (B). aP < 0.05, bP < 0.01 vs model group.
Effects of ACF on PINP and PIIINP contents in rat serum

The serum PINP and PIIINP contents in ACF-2 and -3 group decreased (P < 0.05, P < 0.01) compared to those in model group (Figure 1B).

Expression of MMP-1 mRNA in rat hepatic tissue

There was no significant difference in expression of MMP-1 mRNA between normal group and model group. But the expression of MMP-1 mRNA in ACF groups tended to increase. In ACF-2 and -3 group, the expression of MMP-1 mRNA increased significantly as compared to that in model group (Figure 2).

Figure 2
Figure 2 Analysis of MMP-1 mRNA expression in rat hepatic tissue. M: molecular weight markers (bp); lane 1: normal group; lane 2: model group; lane 3: colchicine group; lane 4: ACF-1 group; lane 5: ACF-2 group; lane 6: ACF-3 group; n = 4.aP < 0.05, bP < 0.01 vs model group.
TIMP-1 protein expression in rat hepatic tissue

The positive expressions of TIMP-1 protein were observed in all groups. In normal group, the mild expression was presented in vascular endothelial cells and sinusoidal space. However, in model group the strong expression was noted around lobule median vein and in portal areas with a positive ratio of 8.84%. After the rats were treated with ACF, TIMP-1 protein expression decreased significantly (P < 0.01, Figure 3).

Figure 3
Figure 3 Immunohistochemical expression of TIMP-1 protein in rat hepatic tissue. (1) The positive ratio detected by Jieda 501 image-analysis system was used as the total protein expression. (2) TIMP-1 is shown as yellow grains in hepatic tissue (immunohistochemistry, ×200). A: normal group; B: model group; C: colchicine group; D: ACF-1 group; E: ACF-2 group; F: ACF-1 group; n = 5. aP < 0.05, bP < 0.01 vs model group.
DISCUSSION

Following chronic liver injury of any etiology, there is progressive fibrosis. The fibrotic process recognizes the involvement of various cells and different factors, leading to excessive fibrogenesis with disruption of intercellular contacts and interactions and of extracellular matrix composition[4-6]. However, hepatic stellate cells (HSCs), together with recruited mononuclear cells and Kupffer cells are by far the key players in liver fibrosis[7-10]. Their cross-talk is triggered and favored by a series of chemical mediators, with a prominent role played by the transforming growth factor beta. Both expression and synthesis of this inflammatory and pro-fibrogenic cytokine are mainly modulated through redox-sensitive reactions. Furthermore, involvement of reactive oxygen species and lipid peroxidation products can be clearly demonstrated in other fundamental events of hepatic fibrogenesis, like activation and effects of stellate cells, expression of metalloproteinases and their specific inhibitors[11-14]. Therefore, anti-oxidative stress can inhibit the process of fibrosis.

Liver fibrosis is characterized by the deposition of excess extracellular matrix (particularly collagen types I and III) in the liver. The aminoterminal procollagen type-I peptide and type-III peptide, which release during collagen types I and III deposition, have been suggested as serum markers of fibrogenesis[15-17]. Their concentrations in serum reflect the ongoing collagen formation in the liver. Extracellular degradation of matrix protein is regulated by MMPs produced by HSCs, which in turn are regulated by several mechanisms including regulation at level of the gene (transcription and proenzyme synthesis), cleavage of the proenzyme to an active form and specific inhibition of activated forms by TIMPs[18].

Extracellular degradation of matrix proteins is regulated by a family of enzymes known as MMPs[19-21]-types I, II, and III collagenases degrade interstitial collagens. Type IV collagenases/gelatinases degrade basement membrane collagen and gelatins and stromelysins which degrade a broad range of substrates including proteoglycans, laminin, gelatins, and fibronectin. However, the activated forms of MMPs in turn are inhibited specifically by TIMPs, such as TIMP-1 and -2[22-26].

Based on this pathophysiological mechanism, therapeutic methods are developed to diminish collagen accumulation, stimulate the collagenolytic process, and/or inhibit stellate cell proliferation[27,28]. However, current therapies targeting at arresting or reversing liver fibrosis are largely ineffective and some have unacceptable side effects in long-term therapy. Recent clinical and experimental observations have demonstrated that Chinese medicines might be of some preventive and therapeutic values against fibrosis[29,30]. SA is a traditional Chinese medicine containing flavonoids, organic acids and triterpenoids, etc.[31]. It can protect animal livers against injury by CCl4 and cirrhosis caused by DMN[1,2].

The present results indicate that ACF decreases the serum content of N-propeptide of types I and III procollagen, elevates the activity of SOD and reduces the MDA content in DMN-treated rat hepatic tissue. In addition, the protein expression of TIMP-1 reduces and the MMP-1 mRNA expression increases in liver fibrosis rats with induced by DMN. The antifibrotic effect of ACF is associated with its influence on lipid peroxidation and collagen synthesis and degradation.

Footnotes

Science Editor Wang XL and Guo SY Language Editor Elsevier HK

References
1.  Liu CY, Gu ZL, Han R, Zu LJ, Zhou WX, Guo CY. Protective effect of extract in seed of Astragalus complanatus on acute hepatic injury induced by carbon tetrachloride in mice. Zhongcaoyao. 2002;33:1104-1106.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Liu CY, Gu ZL, Zhang KP, Zhou WX, Guo CY. Effect of flavonoids of astragali complanali on liver libroses induced by DMN in rats. Zhonggu Yaolixue Tongbao. 2004;20:110-114.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Li ZJ, Jin SHG. Effect of Traditional Chinese Herb Panax Notogenseng (Burk) F.H. Chen on dimethylnitrosamine-induced liver injury in rats. Chengdu Zhongyiyao Daxue Xuebao. 2000;23:31-35.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Gerling B, Becker M, Waldschmidt J, Rehmann M, Schuppan D. Elevated serum aminoterminal procollagen type-III-peptide parallels collagen accumulation in rats with secondary biliary fibrosis. J Hepatol. 1996;25:79-84.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 31]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
5.  Annoni G, Colombo M, Cantaluppi MC, Khlat B, Lampertico P, Rojkind M. Serum type III procollagen peptide and laminin (Lam-P1) detect alcoholic hepatitis in chronic alcohol abusers. Hepatology. 1989;9:693-697.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 46]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
6.  Guéchot J, Laudat A, Loria A, Serfaty L, Poupon R, Giboudeau J. Diagnostic accuracy of hyaluronan and type III procollagen amino-terminal peptide serum assays as markers of liver fibrosis in chronic viral hepatitis C evaluated by ROC curve analysis. Clin Chem. 1996;42:558-563.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Senoo H. Structure and function of hepatic stellate cells. Med Electron Microsc. 2004;37:3-15.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 145]  [Cited by in F6Publishing: 123]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
8.  Safadi R, Friedman SL. Hepatic fibrosis--role of hepatic stellate cell activation. MedGenMed. 2002;4:27.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Marra F, Pinzani M. Role of hepatic stellate cells in the pathogenesis of portal hypertension. Nefrologia. 2002;22 Suppl 5:34-40.  [PubMed]  [DOI]  [Cited in This Article: ]
10.  Reeves HL, Friedman SL. Activation of hepatic stellate cells--a key issue in liver fibrosis. Front Biosci. 2002;7:d808-d826.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 319]  [Cited by in F6Publishing: 353]  [Article Influence: 16.0]  [Reference Citation Analysis (0)]
11.  Zhang X, Yu WP, Gaol , Wei KB, Ju JL, Xu JZ. Effects of lipopolysaecharides stimulated Kupffer cells on activation of rat hepatic stellate cells. World J Gastroenterol. 2004;15:610-613.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Poli G. Pathogenesis of liver fibrosis: role of oxidative stress. Mol Aspects Med. 2000;21:49-98.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 457]  [Cited by in F6Publishing: 472]  [Article Influence: 19.7]  [Reference Citation Analysis (0)]
13.  Parola M, Robino G. Oxidative stress-related molecules and liver fibrosis. J Hepatol. 2001;35:297-306.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 541]  [Cited by in F6Publishing: 539]  [Article Influence: 23.4]  [Reference Citation Analysis (1)]
14.  Gebhardt R. Oxidative stress, plant-derived antioxidants and liver fibrosis. Planta Med. 2002;68:289-296.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 84]  [Cited by in F6Publishing: 76]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
15.  Refik Mas M, Comert B, Oncu K, Vural SA, Akay C, Tasci I, Ozkomur E, Serdar M, Mas N, Alcigir G. The effect of taurine treatment on oxidative stress in experimental liver fibrosis. Hepatol Res. 2004;28:207-215.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 44]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
16.  Xie SB, Yao JL, Zheng RQ, Peng XM, Gao ZL. Serum hyaluronic acid, procollagen type III and IV in histological diagnosis of liver fibrosis. Hepatobiliary Pancreat Dis Int. 2003;2:69-72.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Stone PJ. Potential use of collagen and elastin degradation markers for monitoring liver fibrosis in schistosomiasis. Acta Trop. 2000;77:97-99.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 26]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
18.  Oh S, Afdhal NH. Hepatic fibrosis: are any of the serum markers useful? Curr Gastroenterol Rep. 2001;3:12-18.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 36]  [Cited by in F6Publishing: 38]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
19.  McCrudden R, Iredale JP. Liver fibrosis, the hepatic stellate cell and tissue inhibitors of metalloproteinases. Histol Histopathol. 2000;15:1159-1168.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Okazaki I, Watanabe T, Hozawa S, Arai M, Maruyama K. Molecular mechanism of the reversibility of hepatic fibrosis: with special reference to the role of matrix metalloproteinases. J Gastroenterol Hepatol. 2000;15 Suppl:D26-D32.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 64]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
21.  Benyon RC, Arthur MJ. Extracellular matrix degradation and the role of hepatic stellate cells. Semin Liver Dis. 2001;21:373-384.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 358]  [Cited by in F6Publishing: 388]  [Article Influence: 16.9]  [Reference Citation Analysis (0)]
22.  Schuppan D, Ruehl M, Somasundaram R, Hahn EG. Matrix as a modulator of hepatic fibrogenesis. Semin Liver Dis. 2001;21:351-372.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 391]  [Cited by in F6Publishing: 365]  [Article Influence: 15.9]  [Reference Citation Analysis (0)]
23.  Okazaki I, Arai M, Wada N, Maruyama K. Gene expression of MMPs and TIMPs in the process of hepatic fibrosis. Nihon Rinsho. 1993;51:428-434.  [PubMed]  [DOI]  [Cited in This Article: ]
24.  Murphy FR, Issa R, Zhou X, Ratnarajah S, Nagase H, Arthur MJ, Benyon C, Iredale JP. Inhibition of apoptosis of activated hepatic stellate cells by tissue inhibitor of metalloproteinase-1 is mediated via effects on matrix metalloproteinase inhibition: implications for reversibility of liver fibrosis. J Biol Chem. 2002;277:11069-11076.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 328]  [Cited by in F6Publishing: 350]  [Article Influence: 15.9]  [Reference Citation Analysis (0)]
25.  Arthur MJ. Fibrogenesis II. Metalloproteinases and their inhibitors in liver fibrosis. Am J Physiol Gastrointest Liver Physiol. 2000;279:245-249.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Arthur MJ, Iredale JP, Mann DA. Tissue inhibitors of metalloproteinases: role in liver fibrosis and alcoholic liver disease. Alcohol Clin Exp Res. 1999;23:940-943.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 25]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
27.  Iredale JP. Tissue inhibitors of metalloproteinases in liver fibrosis. Int J Biochem Cell Biol. 1997;29:43-54.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 123]  [Cited by in F6Publishing: 119]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
28.  Murphy F, Arthur M, Iredale J. Developing strategies for liver fibrosis treatment. Expert Opin Investig Drugs. 2002;11:1575-1585.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 20]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
29.  Shimizu I. Antifibrogenic therapies in chronic HCV infection. Curr Drug Targets Infect Disord. 2001;1:227-240.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 29]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
30.  Luo YJ, Yu JP, Shi ZH, Wang L. Ginkgo biloba extract reverses CCl4-induced liver fibrosis in rats. World J Gastroenterol. 2004;10:1037-1042.  [PubMed]  [DOI]  [Cited in This Article: ]
31.  Stickel F, Brinkhaus B, Krähmer N, Seitz HK, Hahn EG, Schuppan D. Antifibrotic properties of botanicals in chronic liver disease. Hepatogastroenterology. 2002;49:1102-1108.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Tang WX, Dan ZL, Yan HM, Wu CH, Zhang G, Liu M, Li Q, Li SB. Experimental study of effect of Ganyanping on fibrosis in rat livers. World J Gastroenterol. 2003;9:1292-1295.  [PubMed]  [DOI]  [Cited in This Article: ]
33.  Du B, You S. Present situation in preventing and treating liver fibrosis with TCM drugs. J Tradit Chin Med. 2001;21:147-152.  [PubMed]  [DOI]  [Cited in This Article: ]