1,020
Views
31
CrossRef citations to date
0
Altmetric
Research Article

Ginsenoside Rg1 protects primary cultured rat hippocampal neurons from cell apoptosis induced by β-amyloid protein

, , &
Pages 501-507 | Received 11 Apr 2010, Accepted 02 Sep 2010, Published online: 26 Mar 2011

Abstract

Objectives: Estrogen is known to prominently benefit neuronal syndromes and neurodegenerative diseases. Ginsenoside Rg1, an active ingredient found in a Chinese plant, ginseng root, was previously demonstrated to exert estrogen-like activity. This study was performed to assess the neuroprotective effect of ginsenoside Rg1 against apoptosis induced by β-amyloid protein 25–35 (Aβ25–35) in primary cultured rat hippocampal neuronal cells as well as in the underlying mechanisms.

Methods: We first measured cell viability and lactate dehydrogenase (LDH) release from primary cultured rat hippocampal neurons. After that, the inhibition effects of ginsenoside Rg1 on neuronal cell apoptosis were evaluated with flow cytometric analysis. Furthermore, western blot analysis was used for detecting the expression of apoptosis-related proteins Bcl-2, Bax, and active caspase 3.

Results: The results show that ginsenoside Rg1 could increase neuronal viability and reduce LDH release; rescue cell apoptosis induced by Aβ25–35; decrease the expression of caspase 3, increase the ratio of Bcl-2/Bax at the protein levels compared with the cells only treated with Aβ25–35.

Conclusions: Taken together, our results indicate that the apoptosis induced by Aβ25–35 could be reversed by ginsenoside Rg1. Furthermore, this neuroprotective effect is probably mediated by up-regulating the ratio of Bcl-2/Bax that activates caspase 3.

Introduction

Pathologically, Alzheimer’s disease (AD) is manifested by selective oxidative stress-induced neuronal cell death, the deposition of β-amyloid (Aβ) peptide into senile plaques in the extracellular space, and the formation of neurofibrillary tangles inside the neuron. Neuronal apoptosis induced by Aβ was proved to be closely associated with AD (CitationKawasumi et al., 2002; CitationPereira et al., 2004). Different therapeutic efforts are targeted at blocking Aβ aggregation, Aβ production, Aβ-induced toxicity, and oxidative stress (CitationZhu et al., 2004; CitationIrie et al., 2005; CitationTanzi & Bertram, 2005). Estrogen is shown to decrease the production of Aβ and to be effective against Aβ-induced toxicity in neurons (CitationGreen et al., 1996; CitationXu et al., 1998; CitationGoodenough et al., 2003). However, we have to recognize the side effects of estrogen replacement therapy (ERT). It probably increases proliferative and oncogenic activities on non-neuronal cells such as cells in the breast and endometrium (CitationBeresford et al., 1997; CitationCollaborative Group on Hormonal Factors in Breast Cancer, 1997).

Accordingly, growing attention has been paid to phytoestrogens for their potential therapeutic use in a range of hormone-dependent diseases. They have fewer side effects and have been proven to possess neuroprotective activity (CitationMackey & Eden, 1998; CitationKim et al., 2005). Panax ginseng C.A. Meyer (Araliaceae) is an important Chinese herbal medicine that has been widely used in restoring harmony and improving memory for thousands of years in China, and recently it is being widely studied in the west. Ginsenosides were reported to be the most abundant and active compounds contributing to the pharmacological actions of ginseng (CitationHuang & Williams, 1999). As one of nearly 40 different ginsenosides that have been identified from ginseng, ginsenoside Rg1 has been attracted considerable attention. In recent years, ginsenoside Rg1 has been termed as a phytoestrogen as it exhibits estrogenic properties (CitationChan et al., 2002; CitationLeung et al., 2007; CitationGao et al., 2009). However, the molecular mechanisms underlying the neuroprotective effect of ginsenoside Rg1 have not been fully studied. Our experiments were performed to explore whether ginsenoside Rg1 could protect primary cultured rat neuronal cells against insult of Aβ25–35. Furthermore, in our experiments, it was demonstrated that ginsenoside Rg1 regulates the expression of Bcl-2, bax, and caspase 3 and thus inhibit the neuronal cell apoptosis induced by Aβ.

Materials and methods

General chemicals

Ginsenoside Rg1 was purchased from the National Institute for the Control of Pharmaceutical and Biological Products (Beijing, China). The purity of ginsenoside Rg1 for experiments was >99%. Aβ25–35, 17β-estradiol (E2), dimethyl sulfoxide (DMSO), arabinosylcytosine, 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-tetrazolium bromide (MTT) and poly-d-lysine were purchased from Sigma-Aldrich (St. Louis, MO).

Primary antibodies of Bcl-2, Bax, and active form of caspase 3 were purchased from Cell Signaling Technology (San Francisco, CA). Dulbecco’s modified Eagle’s medium/F12 medium (DMEM/F12) and trypsin were purchased from Gibco (Burlington, Ontario, Canada). Fetal bovine serum was obtained from Sijiqing (Hangzhou, China).

Primary neuronal cell cultures

Primary cultured hippocampal neurons were prepared from embryonic brains of Wistar rats at Days 16–18 of gestation as previously described (CitationBastianetto et al., 1999; CitationXie et al., 2000). After removal of meninges in DMEM/F12 medium, hippocampi were cut into small pieces of 1 mm3 and treated with 0.125% trypsin for 15 min at 37°C. The reaction was stopped using DMEM/F12 medium supplemented with 10% fetal bovine serum. The pellet was dissociated by repeated passage through a series of fire-polished constricted Pasteur pipettes and passed through a nylon mesh. Finally, cells were plated on polyethylenimine-coated (0.1 mg/mL) 6-well, 12-well, 24-well plates or culture flasks with 90% DMEM/F12 medium, 10% fetal bovine serum, 10 U/mL penicillin and 10 U/mL streptomycin at a density of 1 × 105 cells/mL. The cultures were placed in a humidified atmosphere of 95% air and 5% CO2 at 37°C. The cells were exposed to arabinosylcytosine (5 μg/mL) for 24 h at Day 3 to suppress the proliferation of glial cells. Cells were used for experiments after 10 days in vitro.

Exposure to neuronal cells

Ginsenoside Rg1 was dissolved in sterile distilled water at a concentration of 0.4 mg/mL as a stock solution. Aβ25–35 at a concentration of 1 mg/mL was dissolved in DMEM/F12 medium and incubated for 7 days at 37°C prior to use. E2 was dissolved at 10 mM in ethanol and then diluted into DMEM/F12 media to a concentration of 10 μM.

Neuronal cells were treated with Aβ25–35 and either ginsenoside Rg1 or E2 after refreshing the medium at Day 7. The experiments were performed 72 h later.

Analysis of cell injury

Cell viability was determined by the MTT method. Cells (5 × 104 cells/well) were seeded in 96-well plates. In brief, after various treatments, cells were incubated with MTT solution (25 μL, 5 mg/mL). Four hours later, cells were exposed to DMSO by shaking for 10 min. Absorbance of each well was measured at a wavelength of 570 nm with a universal microplate reader (ELx 800, BioTek Instruments, Winooski, VT). The optical density of untreated controls was set at 100%.

Cell viability was also evaluated by measurement of lactate dehydrogenase (LDH) (Biovision, USA) released into the medium. LDH activity was determined according to the manufacturer’s instructions. Colorimetric absorbance was measured at 570 nm with a microplate reader (ELx800).

Assays were repeated in three independent experiments, each performed in quintuplicate.

Flow cytometric analysis of Annexin V/PI staining

Cell apoptosis was investigated by staining cells with both Annexin V and PI (Biovision, Mountain View, CA). The 1–5 × 105 cells were collected by centrifugation after various treatments. Then cells were incubated with Annexin V and PI according to the manufacturer’s instructions for 5 min at room temperature in the dark. Apoptotic cells can be recognized and distinguished from necrotic cells using flow cytometric analysis (FACSCalibur, Franklin Lakes, NJ). The percentage of apoptotic cells was determined from a DNA histogram as the ratio of the hypo-diploid cell population (the sub-G1 peak) to the total number of cells.

Protein expression detection by western blot analysis

The protein expression of Bcl-2, Bax, and active caspase 3 was determined by western blot analysis. Cells were collected on ice in lysis buffer containing Tris–HCl (pH 6.8, 125 mM), 2% sodium dodecyl sulfate (SDS), 10% glycerol, and 2-mercaptoethanol (pH 6.8, 200 mM). Then SDS-polyacrylamide gel electrophoresis was performed and the proteins were transferred to a polyvinylidene fluoride (PVDF) membrane. The membrane was washed in Tris-buffered saline and blocked for 2 h in 5% nonfat milk in Tris-buffered saline containing 0.1% Tween 20 (TBS/T) at room temperature. The membrane was incubated in 5% milk with primary antibodies (1:1000) overnight at 4°C. After that, the membrane was washed three times for 5 min each with 10 mL TBS/T and incubated with secondary anti-rabbit HRP antibody (1:3000) with gentle agitation for 1.5 h at room temperature. The bands were visualized by the enhanced chemiluminescence (ECL) method and exposed to an X-ray film. The intensity of the bands was quantified using a densitometer. For quantification, densities were normalized by corresponding blots for actin.

Statistical analysis

All data are expressed as mean ± SD. Statistical comparisons were performed by one-way analysis of variance (ANOVA) and the least significant difference (LSD) t-test. A P-value of <0.05 was considered statistically significant.

Results

Establishment of Aβ25–35-insulted primary culture neuronal cell model

The aggregation of Aβ and Aβ25–35 are considered to be critical events that induce neuronal apoptosis (CitationPike et al., 1995). Therefore, in this study we used Aβ25–35 to induce neuronal toxicity. Various concentrations of Aβ25–35 were added into the culture medium to determine a suitable concentration for the model. Cell viability was assessed by MTT method and LDH release assessment. As shown in , the toxicity of Aβ25–35 to neuronal cells was in a dose-dependent manner. The results showed decreased levels of MTT reduction by 34.7% after treatment with 5 μM Aβ25–35 and by 30% after treatment with 10 μM Aβ25–35 for 3 days at 37°C (); LDH release assessment obtained the same results (). Accordingly, we eventually chose a 3-day treatment of 5 μM Aβ25–35 for further experiments.

Figure 1.  Effect of Aβ25–35 on the viability of primary cultured neuronal cells. The 7-day-old cells were treated with various concentrations of Aβ25–35 (0, 2.5, 5, 7.5, 10 μm) for 3 days and then measured. (A) Cell viability and (B) LDH release. Data represent means ± SD, n = 3.

Figure 1.  Effect of Aβ25–35 on the viability of primary cultured neuronal cells. The 7-day-old cells were treated with various concentrations of Aβ25–35 (0, 2.5, 5, 7.5, 10 μm) for 3 days and then measured. (A) Cell viability and (B) LDH release. Data represent means ± SD, n = 3.

Ginsenoside Rg1 inhibits the cytotoxicity of Aβ25–35

The neuroprotective effects of ginsenoside Rg1 in preventing Aβ25–35-induced cytotoxicity were determined. Neuronal cells underwent various treatments for 3 days before the viability test. As illustrated in , after treatment with Aβ25–35, the cell viability was decreased remarkably to 65.3 ± 5.5% and LDH release was increased to 293.7 ± 19.3%. The cell viability was decreased and LDH release was inhibited in the ginsenoside Rg1 group and E2 group compared with the model group. Ginsenoside Rg1 groups produced dose-dependent effects on LDH release.

Figure 2.  Ginsenoside Rg1 inhibits Aβ25–35-induced neurotoxicity of primary cultured neuronal cells. Primary cultured neuronal cells were treated with E2, various concentrations of ginsenoside Rg1 and Aβ25–35 of 5 μM for 3 days. (A) Cell viability and (B) LDH release. Data represent means ± SD, n = 3. *P < 0.01 versus the control group, **P < 0.05 versus the model group, and #P < 0.01 versus the model group.

Figure 2.  Ginsenoside Rg1 inhibits Aβ25–35-induced neurotoxicity of primary cultured neuronal cells. Primary cultured neuronal cells were treated with E2, various concentrations of ginsenoside Rg1 and Aβ25–35 of 5 μM for 3 days. (A) Cell viability and (B) LDH release. Data represent means ± SD, n = 3. *P < 0.01 versus the control group, **P < 0.05 versus the model group, and #P < 0.01 versus the model group.

Ginsenoside Rg1 protects apoptosis induced by Aβ25–35

To determine the cell apoptosis after various treatments, Annexin V and PI were used to stain cells and were then distinguished by flow cytometric analysis. As seen in , the early apoptotic cells are presented in the lower right quadrant of the fluorescence-activated cell-sorting histogram (Annexin V FITC+/PI). The late apoptotic cells are presented in the upper right quadrant (Annexin V FITC+/PI+). As seen in , various concentrations of ginsenoside Rg1 significantly reduced the numbers of early apoptotic cells to 11.5 ± 1.32% (10−8 M), 7.12 ± 1.16% (10−7 M), 6.88 ± 2.26% (10−6 M) compared with 18.19 ± 0.77% of the model group. The numbers of early apoptotic cells were reduced to 6.56 ± 1.91% in the E2 group. These results suggest that both ginsenoside Rg1 and E2 could rescue apoptosis induced by Aβ25–35.

Figure 3.  Detection of Ginsenoside Rg1 protecting Aβ25–35-induced apoptosis by flow cytometry analysis of Annexin V/PI staining. The early apoptotic cells are presented in the lower right quadrant (Annexin V FITC+/PI). The late apoptotic cells are presented in the upper right quadrant (Annexin V FITC+/PI+). Viable cells are presented in the lower left quadrant (Annexin V FITC/PI). Necrotic cells are shown in the upper left quadrants (Annexin V FITC/PI+). (A) Control, (B) model, (C) Rg1 10−8 M, (D) Rg1 10−7 M, (E) Rg1 10−6 M, and (F) E2 10−8 M.

Figure 3.  Detection of Ginsenoside Rg1 protecting Aβ25–35-induced apoptosis by flow cytometry analysis of Annexin V/PI staining. The early apoptotic cells are presented in the lower right quadrant (Annexin V FITC+/PI−). The late apoptotic cells are presented in the upper right quadrant (Annexin V FITC+/PI+). Viable cells are presented in the lower left quadrant (Annexin V FITC−/PI−). Necrotic cells are shown in the upper left quadrants (Annexin V FITC−/PI+). (A) Control, (B) model, (C) Rg1 10−8 M, (D) Rg1 10−7 M, (E) Rg1 10−6 M, and (F) E2 10−8 M.

Figure 4.  The percentage of apoptotic cells after Annexin V/PI staining. Data represent means ± SD, n = 3. *P <0.01 versus the control group and **P < 0.05 versus the model group.

Figure 4.  The percentage of apoptotic cells after Annexin V/PI staining. Data represent means ± SD, n = 3. *P <0.01 versus the control group and **P < 0.05 versus the model group.

Expression of Bcl-2, Bax, and active caspase 3 protein

As shown in and , levels of Bcl-2 protein were up-regulated after being treated with ginsenoside Rg1 and E2, and levels of Bax protein were down-regulated ( and ). It resulted in an up-regulation of the Bcl-2/Bax ratio ( and ). As illustrated in and , ginsenoside Rg1 down-regulated the activation of caspase 3 similar to E2.

Figure 5.  Protein levels of Bcl-2, Bax, Bcl-2/Bax, and active caspase 3 were changed by ginsenoside Rg1.

Figure 5.  Protein levels of Bcl-2, Bax, Bcl-2/Bax, and active caspase 3 were changed by ginsenoside Rg1.

Figure 6.  The ratio of protein expressions. (A) Bcl-2/actin, (B) Bax/actin, (C) Bcl-2/Bax, and (D) active caspase 3/actin. Data represent means ± SD. n = 3. *P < 0.01 versus control group, **P < 0.05 versus the model group, and #P <0.01 versus the model group.

Figure 6.  The ratio of protein expressions. (A) Bcl-2/actin, (B) Bax/actin, (C) Bcl-2/Bax, and (D) active caspase 3/actin. Data represent means ± SD. n = 3. *P < 0.01 versus control group, **P < 0.05 versus the model group, and #P <0.01 versus the model group.

Discussion

A large body of evidence suggests that estrogen plays a profound neuroprotective role in Alzheimer’s and Parkinson’s diseases (CitationBehl, 2002). In recent years, because of the side effects of estrogen in clinical evidence, ginseng has been widely studied as it was shown to possess estrogen-like activity. It was found that ginsenoside Rg1 pretreatment reduced inducible nitric oxide (NO) synthase protein level and NO production (CitationChen et al., 2003). CitationGao et al. (2009) have reported that ginsenoside Rg1 may attenuate 6-OHDA-induced apoptosis and its action might involve the activation of IGF-IR signaling pathway and ER signaling pathway. CitationLeung et al. (2007) have reported that ginsenoside Rg1 inhibits the mitochondrial apoptotic pathway and increases the survival chance of primary cultured nigral neurons against rotenone toxicity. However, its molecular mechanisms implicated in the protective actions are not fully understood. In the present study, we attempt to study the mechanism of the neuroprotective effect of ginsenoside Rg1, one of the most abundant ginsenosides found in ginseng.

Our results show that ginsenoside Rg1 could protect neurons from damage induced by Aβ25–35 insults similar to E2. Additionally, ginsenoside Rg1 increased neuronal viability and reduced LDH release in cultured neuronal cells. We found that after being treated with Aβ25–35 for 3 days, cell viability had dramatically decreased and that ginsenoside Rg1 could reverse this. Moreover, in Annexin V/PI staining tests, with ginsenoside Rg1 treatment, the neuronal cells appeared less apoptotic than those treated with only Aβ25–35. All these findings demonstrated that ginsenoside Rg1 could protect cells against Aβ25–35-induced toxicity in primary cultured neurons.

Caspases (cysteine aspartate-specific proteases) are a family of intracellular proteins involved in the initiation and execution of apoptosis (CitationAshe & Berry, 2003). Apoptosis is promoted by proteolysis that is largely due to the activity of caspases (CitationMiller, 1997; CitationTakai et al., 1998). Caspases are synthesized as inactive proenzymes in the cytoplasm and are activated by cleavage of specified aspartate residues, thus initiating a cascade of proteolytic cleavage leading to activation of downstream caspases with cellular substrates, such as poly(ADP-ribose) polymerase (PARP) and lamins (CitationJellinger, 2000). To date, >14 mammalian caspases have been identified (CitationThornberry & Lazebnik, 1998). Caspase 3 is a key executor caspase (CitationKim et al., 1997); its activation is considered specific for the apoptotic process and defines an irreversible stage in the cell-death cascade (CitationJellinger, 2000). In our experiment, ginsenoside Rg1 and E2 decreased the expression of active caspase 3 at the protein levels compared with the cells treated with Aβ25–35. This indicated that the activation of caspase 3 plays an important role in the neuroprotective effect of both ginsenoside Rg1 and E2.

It was found that in apoptosis triggered by many stimuli, mitochondria play a pivotal role in coordinating caspase activation through the release of cytochrome c (Solange & Martinou, 2000). Bcl-2 is an integral membrane protein located mainly on the outer membrane of mitochondria, as well as a gene with clear anti-apoptotic properties in neurodegenerative conditions (CitationYang et al., 1997; CitationNuydens et al., 2000). Thus, one possible role of Bcl-2 in the prevention of apoptosis is to block cytochrome c release from mitochondria (CitationYang et al., 1997). Currently, 15 Bcl-2 family proteins have been identified in mammals (CitationGross et al., 1999). This family includes proteins that can promote either cell survival, such as Bcl-2, Bcl-XL, Mcl-1, A1, Bcl-W, or cell death like Bax, Bak, Bcl-XS, Bok (CitationAdams & Cory, 1998; CitationChao & Korsmeyer, 1998). CitationSato et al. (1994) suggested that Bax is a cell-death effector whose activity is neutralized by binding of Bcl-2. The ratio between these two subsets helps determine, in part, the susceptibility of cells to a death signal (CitationOltvai et al., 1993). Data from the study demonstrated an increase of Bcl-2, decrease of Bax, as well as a dramatic up-regulation of the Bcl-2/Bax ratio in the ginsenoside Rg1 group and E2 group neuron cells. These results are in line with our other results described earlier. This indicated that ginsenoside Rg1 and E2 protect the neuron cells from apoptosis induced by Aβ25–35 insults, probably because it could increase the ratio of Bcl-2/Bax that down-regulates the activation of caspase 3 and finally inhibits neuronal cell apoptosis. Moreover, further studies are required to determine whether estrogen receptors are involved in ginsenoside Rg1-related neuroprotection, or whether there are other signal transduction pathways involved.

Conclusion

In conclusion, ginsenoside Rg1 exerted a neuroprotective role on the apoptosis neurons induced by Aβ25–35 similar to E2. However, there are fewer side effects of using ginsenoside Rg1 compared with E2. This proves that more attention should be paid to clarify the further mechanisms of ginsenoside Rg1, which will help us to discover more potent neuroprotective candidates.

Declaration of interest

This work was supported by the Health Department of Shandong Province, Jinan, China. The authors report no conflicts of interest. The authors alone are responsible for the content and writing of the paper.

References

  • Adams JM, Cory S. (1998). The Bcl-2 protein family: arbiters of cell survival. Science 281:1322–1326.
  • Ashe PC, Berry MD. (2003). Apoptotic signaling cascades. Prog Neuropsychopharmacol Biol Psychiatry 27:199–214.
  • Bastianetto S, Ramassamy C, Poirier J, Quirion R. (1999). Dehydroepiandrosterone (DHEA) protects hippocampal cells from oxidative stress-induced damage. Brain Res Mol Brain Res 66:35–41.
  • Behl C. (2002). Oestrogen as a neuroprotective hormone. Nat Rev Neurosci 3:433–442.
  • Beresford SA, Weiss NS, Voigt LF, McKnight B. (1997). Risk of endometrial cancer in relation to use of oestrogen combined with cyclic progestagen therapy in postmenopausal women. Lancet 349:458–461.
  • Chan RY, Chen WF, Dong A, Guo D, Wong MS. (2002). Estrogen-like activity of ginsenoside Rg1 derived from Panax notoginseng. J Clin Endocrinol Metab 87:3691–3695.
  • Chao DT, Korsmeyer SJ. (1998). BCL-2 family: regulators of cell death. Annu Rev Immunol 16:395–419.
  • Chen XC, Zhu YG, Zhu LA, Huang C, Chen Y, Chen LM, Fang F, Zhou YC, Zhao CH. (2003). Ginsenoside Rg1 attenuates dopamine-induced apoptosis in PC12 cells by suppressing oxidative stress. Eur J Pharmacol 473:1–7.
  • Collaborative Group on Hormonal Factors in Breast Cancer. (1997). Breast cancer and hormone replacement therapy: collaborative reanalysis of data from 51 epidemiological studies of 52,705 women with breast cancer and 108,411 women without breast cancer. Lancet 350:1047–1059.
  • Desagher S, Martinou JC. (2000). Mitochondria as the central control point of apoptosis. Trends Cell Biol 10:369–377.
  • Gao QG, Chen WF, Xie JX, Wong MS. (2009). Ginsenoside Rg1 protects against 6-OHDA-induced neurotoxicity in neuroblastoma SK-N-SH cells via IGF-I receptor and estrogen receptor pathways. J Neurochem 109:1338–1347.
  • Goodenough S, Schäfer M, Behl C. (2003). Estrogen-induced cell signalling in a cellular model of Alzheimer’s disease. J Steroid Biochem Mol Biol 84:301–305.
  • Green PS, Gridley KE, Simpkins JW. (1996). Estradiol protects against beta-amyloid (25–35)-induced toxicity in SK-N-SH human neuroblastoma cells. Neurosci Lett 218:165–168.
  • Gross A, McDonnell JM, Korsmeyer SJ. (1999). BCL-2 family members and the mitochondria in apoptosis. Genes Dev 13:1899–1911.
  • Huang KC, Williams WM. (1999). The Pharmacology of Chinese Herbs. Boca Raton: CRC Press.
  • Irie K, Murakami K, Masuda Y, Morimoto A, Ohigashi H, Ohashi R, Takegoshi K, Nagao M, Shimizu T, Shirasawa T. (2005). Structure of beta-amyloid fibrils and its relevance to their neurotoxicity: implications for the pathogenesis of Alzheimer’s disease. J Biosci Bioeng 99:437–447.
  • Jellinger KA. (2000). Cell death mechanisms in Parkinson’s disease. J Neural Transm 107:1–29.
  • Kawasumi M, Hashimoto Y, Chiba T, Kanekura K, Yamagishi Y, Ishizaka M, Tajima H, Niikura T, Nishimoto I. (2002). Molecular mechanisms for neuronal cell death by Alzheimer’s amyloid precursor protein-relevant insults. Neurosignals 11:236–250.
  • Kim H, Park BS, Lee KG, Choi CY, Jang SS, Kim YH, Lee SE. (2005). Effects of naturally occurring compounds on fibril formation and oxidative stress of beta-amyloid. J Agric Food Chem 53:8537–8541.
  • Kim TW, Pettingell WH, Jung YK, Kovacs DM, Tanzi RE. (1997). Alternative cleavage of Alzheimer-associated presenilins during apoptosis by a caspase-3 family protease. Science 277:373–376.
  • Leung KW, Yung KK, Mak NK, Chan YS, Fan TP, Wong RN. (2007). Neuroprotective effects of ginsenoside-Rg1 in primary nigral neurons against rotenone toxicity. Neuropharmacology 52:827–835.
  • Mackey R, Eden J. (1998). Phytoestrogens and the menopause. Climacteric 1:302–308.
  • Miller DK. (1997). The role of the caspase family of cysteine proteases in apoptosis. Semin Immunol 9:35–49.
  • Nuydens R, Dispersyn G, Van Den Keiboom G, de Jong M, Connors R, Ramaekers F, Borgers M, Geerts H. (2000). Bcl-2 protects against apoptosis-related microtubule alterations in neuronal cells. Apoptosis 5:43–51.
  • Oltvai ZN, Milliman CL, Korsmeyer SJ. (1993). Bcl-2 heterodimerizes in vivo with a conserved homolog, Bax, that accelerates programmed cell death. Cell 74:609–619.
  • Pereira C, Ferreiro E, Cardoso SM, de Oliveira CR. (2004). Cell degeneration induced by amyloid-beta peptides: implications for Alzheimer’s disease. J Mol Neurosci 23:97–104.
  • Pike CJ, Walencewicz-Wasserman AJ, Kosmoski J, Cribbs DH, Glabe CG, Cotman CW. (1995). Structure-activity analyses of beta-amyloid peptides: contributions of the beta 25–35 region to aggregation and neurotoxicity. J Neurochem 64:253–265.
  • Sato T, Hanada M, Bodrug S, Irie S, Iwama N, Boise LH, Thompson CB, Golemis E, Fong L, Wang HG. (1994). Interactions among members of the Bcl-2 protein family analyzed with a yeast two-hybrid system. Proc Natl Acad Sci USA 91:9238–9242.
  • Takai N, Nakanishi H, Tanabe K, Nishioku T, Sugiyama T, Fujiwara M, Yamamoto K. (1998). Involvement of caspase-like proteinases in apoptosis of neuronal PC12 cells and primary cultured microglia induced by 6-hydroxydopamine. J Neurosci Res 54:214–222.
  • Tanzi RE, Bertram L. (2005). Twenty years of the Alzheimer’s disease amyloid hypothesis: A genetic perspective. Cell 120:545–555.
  • Thornberry NA, Lazebnik Y. (1998). Caspases: Enemies within. Science 281:1312–1316.
  • Xie C, Markesbery WR, Lovell MA. (2000). Survival of hippocampal and cortical neurons in a mixture of MEM+ and B27-supplemented neurobasal medium. Free Radic Biol Med 28:665–672.
  • Xu H, Gouras GK, Greenfield JP, Vincent B, Naslund J, Mazzarelli L, Fried G, Jovanovic JN, Seeger M, Relkin NR, Liao F, Checler F, Buxbaum JD, Chait BT, Thinakaran G, Sisodia SS, Wang R, Greengard P, Gandy S. (1998). Estrogen reduces neuronal generation of Alzheimer beta-amyloid peptides. Nat Med 4:447–451.
  • Yang J, Liu X, Bhalla K, Kim CN, Ibrado AM, Cai J, Peng TI, Jones DP, Wang X. (1997). Prevention of apoptosis by Bcl-2: Release of cytochrome c from mitochondria blocked. Science 275:1129–1132.
  • Zhu X, Raina AK, Lee HG, Casadesus G, Smith MA, Perry G. (2004). Oxidative stress signalling in Alzheimer’s disease. Brain Res 1000:32–39.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.