Open Access

Age‑related changes in hypertensive brain damage in the hippocampi of spontaneously hypertensive rats

  • Authors:
    • Yali Li
    • Jian Liu
    • Dengfeng Gao
    • Jin Wei
    • Haifeng Yuan
    • Xiaolin Niu
    • Qiaojun Zhang
  • View Affiliations

  • Published online on: February 3, 2016     https://doi.org/10.3892/mmr.2016.4853
  • Pages: 2552-2560
  • Copyright: © Li et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The aim of the present study was to investigate the age‑related alterations in hypertensive brain damage in the hippocampi of spontaneously hypertensive rats (SHR) and the underlying mechanisms. Aging resulted in a significant increase in the number of activated astrocytes and apoptotic cells in the SHR group, which was accompanied by increased expression of oxidative stress markers (iNOS and gp47phox) and apoptotic regulatory proteins (Bax and caspase‑3). In addition, the expression of PPAR‑γ and Bcl‑2 were progressively reduced with increasing age in the SHR group. The 32 and 64‑week‑old SHRs exhibited significantly increased numbers of apoptotic cells, oxidative stress markers and pro‑apoptotic proteins compared with age‑matched WKY rats, which was accompanied by reduced expression of PPAR‑γ. Compared with the 16 and 32‑week‑old WKY group, the 64‑week‑old WKY rats exhibited increased oxidative stress and pro‑apoptotic markers, and increased levels apoptotic cells. In conclusion, the present study indicated that both aging and hypertension enhanced brain damage and oxidative stress injury in the hippocampi of SHRs, indicated by an increased presence of apoptotic cells and astrocytes. In addition, reduced expression of PPAR‑γ may contribute to the age‑related brain damage in SHRs.

Introduction

Hypertension is a common disease that may result in pathological changes in important organs including the brain, heart and the kidneys. Hypertension-associated brain damage can induce cerebral dysfunction, which may manifest as cognitive decline (1,2). Although it is generally believed that hypertension-associated vascular events are the main causes of dementia, many patients experience cognitive decline prior to a stroke (3). In addition, there is a delay between the onset of hypertension and cerebrovascular disease. During this period, a series of pathological alterations occur in the brain, including oxidative stress and cellular apoptosis (4). Furthermore, hypertension is known to enhance the levels of oxidative stress in the brain and vascular tissues (4). Additionally, increased oxidative stress can lower the level of neuronal activity and more importantly, result in the massive entrance of Ca2+ into neurons, thus activating proapoptotic mechanisms and leading to neuronal loss (5). However, few studies have assessed the age-related changes of hypertensive brain damage in spontaneously hypertensive rats (SHR). Studies investigating the pathophysiological mechanisms involved in hypertension-associated cellular apoptosis and oxidative stress injury may provide new drug targets to prevent the progression of the pathological changes observed in hypertension.

SHRs are the model most extensively used for the evaluation of hypertensive brain damage and the potential treatments. The time-dependent rise in arterial blood pressure and the occurrence of brain atrophy, loss of nerve cells and glial reaction share, to an extent, similarities with the process occurring in human hypertensive brains (6). SHRs, therefore, represent a reasonable model of hypertension-associated brain damage. The hippocampus is a brain region involved in learning and memory (7), and is particularly sensitive to ischemia and represents a predictive site for assessing brain damage (8). However, the majority of investigations on the sensitivity of the hippocampus to hypertension and/or ischemia have been performed in acute conditions or following abrupt elevation of blood pressure (9). Only few studies have evaluated the influence of chronic hypertension or ischemia on the structure of the hippocampus (6). Furthermore, the SHR models used in previous studies were ~6–26-week-old, which is not representative of the advanced hypertensive period (3,10,11). Peroxisome proliferator-activated receptor (PPAR)-γ is a ligand-activated transcription factor belonging to the nuclear hormone receptor superfamily, and is involved in the oxidative stress response (12). Oxidative stress is able to attenuate PPAR-γ expression and activity in vascular endothelial cells through the suppression of PPAR-γ transcription (13). In addition, the PPAR-γ agonist rosiglitazone has been shown to protect QZG cells against oxidative stress injury through modulating the PPAR-γ pathway (14). Thus, PPAR-γ may participate in oxidative stress-related injury.

The present study assessed whether hypertension is able to induce apoptosis and/or necrosis in the hippocampus of SHRs, and the associated age-related alterations. In addition, the oxidative stress pathway was investigated as to whether it was involved in this process and its potential underlying mechanisms.

Materials and methods

Animals

16, 32 and 64-week-old male normotensive Wistar-Kyoto (WKY) rats and SHRs were purchased from Shanghai Laboratory Animal Center (n=6 per age group; Shanghai, China). They were housed under humidity (50–60%), temperature (21–23°C) and light cycle (12 h light/dark) controlled conditions, with free access to food and water. All procedures were conducted according to the guidelines established by the Institutional Animal Care and Use Committee and the National Institutes of Health (Bethesda, MD, USA). The study was approved by the ethics committee of The Second Affiliated Hospital or Xi'ab Jiaotong University (Xi'an, China).

Blood pressure measurements

Systolic blood pressure (SBP) was measured by an indirect tail-cuff method (BP-2000; Visitech Systems, Inc., Apex, NC, USA). Briefly, unanesthetized rats were placed in a holding device mounted on a thermostatically controlled warming plate to make the pulsations of the tail artery detectable. Tail cuffs were fixed on the animals, which were allowed to acclimate to the cuffs for 10 min prior to each pressure recording session.

Brain tissue preparation

Rats were anesthetized with 10% chloral hydrate. For western blot analysis, rats (n=3) were perfused transcardially with 0.9% saline (pH 7.4), following which the brains were rapidly removed and stored in sample protectors (Takara Biotechnology, Inc., Dalian, China) until use. For immunohistochemistry and the terminal deoxynucleotidyl transferase-mediated dUTP end-labeling (TUNEL) assay, rats (n=3) were perfused intraventricularly with 0.9% saline (pH 7.4) and 4% formaldehyde. The brains were subsequently removed and postfixed in 4% formaldehyde overnight. Targeted brain pieces were chosen, dehydrated, embedded in paraffin (Beyotime Institute of Biotechnology, Haimen, China) and cut into 10-µm-thick sections. A total of 3 sections containing both the cortex and hippocampus were selected from each rat for immunohistochemical and TUNEL analysis.

Reverse transcription-quantitative polymerase chain reaction (RT-qPCR)

Total RNA was extracted from the hippocampus with the use of TRIzol reagent (Invitrogen; Thermo Fisher Scientific, Inc., Waltham, MA, USA) and the RNA samples were transcribed to cDNA using a PrimeScript RT Master Mix kit (Takara Biotechnology, Inc.) according to the manufacturer's instructions. RT-qPCR was performed with SYBR ExScript RT-PCR Kit (Takara Biotechnology, Inc.) on an iQ Multicolor Real-Time PCR Detection system (Bio-Rad Laboratories, Inc., Hercules, CA, USA). The primers used were as follows: PPAR-γ, forward 5′-GGAGCCTAAGTTTGAGTTTGCTGTG-3′ and reverse 5′-TGCAGCAGGTTGTCTTGGATG-3′; caspase-3, forward 5′-GAGACAGACAGTGGAACTGACGATG-3′ and reverse 5′-GGCGCAAAGTGACTGGATGA-3′; Bax, forward 5′-GCGTCCACCAAGAAGCTGA-3′ and reverse 5′-ACCACCCTGGTCTTGGATCC-3′; Bcl-2, forward 5′-GACTGAGTACCTGAACCGGCATC-3′ and reverse 5′-CTGAGCAGCGTCTTCAGAGACA-3′; and β-actin, forward 5′-GGAGATTACTGCCCTGGCTCCTA-3′ and reverse 5′-GACTCATCGTACTCCTGCTTGCTG-3′. The primers were designed and synthesized by Takara Biotechnology, Inc. Amplification was performed at 95°C for 30 sec, followed by 40 cycles of 95°C for 3 sec and 60°C for 30 sec. Cycle threshold values were obtained from the Bio-Rad iQ5 2.0 Standard Edition optical System software (Bio-Rad Laboratories, Inc.). Relative quantification was performed using the comparative method (2−ΔΔCq) (15) and data was presented as the mean ± standard deviation of three separate experiments conducted in triplicate.

Immunohistochemistry

In brief, sections were deparaffinized and rehydrated according to standard protocols. Subsequently, sections were treated with 3% H2O2 for 20 min, followed by incubation with 0.3% Triton X-100 for 30 min and 10% goat serum for 1 h at room temperature. The slides were then incubated overnight at 4°C with rabbit polyclonal anti-glial fabrillary acidic protein (GFAP; 1:2,000; ab16997; Abcam, Cambridge, UK). Following washing in phosphate-buffered saline (PBS), the sections were incubated with biotinylated goat anti-rabbit IgG (1:100; A10547; Invitrogen; Thermo Fisher Scientific, Inc.). The immuno-complexes were detected using 3,3′-diaminobenzidine. The series of sections obtained from the different groups were processed in the same way to avoid alterations in the intensity of the staining occurring as a result of different incubation conditions. For the negative control, the primary antibody was replaced with PBS. Activated astrocytes express high levels of GFAP and have altered morphology (larger cell bodies and thick branches). Using Image-Pro Plus version 6.0 (Media Cybernetics, Inc., Rockville, MD, USA) to count activated astrocytes, a threshold was set up for positive cell area, and within these area limits, each nucleated cell exhibiting GFAP labeling was counted. Labeled astrocytes were investigated in the right and left hippocampus in 5–6 sections per animal, using three animals per group.

Western blot analysis

For western blot analysis, the brain tissues were homogenized and the total proteins were extracted using radioimmunoprecipitation assay lysis buffer. The protein concentrations were determined using a bicinchoninic acid assay kit (Beyotime Institute of Biotechnology). A total of 20 µg of each sample was separated on 12% sodium dodecyl sulfate polyacrylamide gels, transferred to polyvinylidene fluoride membranes (both from Beyotime Institute of Biotechnology), and blocked in 10% non-fat milk at room temperature for 2 h. Membranes were incubated overnight at 4°C with rabbit polyclonal antibodies against PPAR-γ (1:400; ab66343; Abcam), rabbit polyclonal antibodies against gp47phox (bs3261), Bax (bs2538), Bcl-2 (bs1511), caspase-3 (bs7004; all 1:800; Bio-World, Dublin, OH, USA) and a mouse monoclonal antibody against inducible nitric oxide synthase (iNOS; 1:500; sc-7271; Santa Cruz Biotechnology, Inc., Dallas, TX, USA). The blots were then washed with Tris-buffered saline-Tween-20 (0.1%) 3 times, and incubated with secondary antibodies horseradish peroxidase (HRP)-conjugated goat anti-rabbit immunoglobulin G (1:5,000; BS13278; Bio-World) and HRP-conjugated goat anti-mouse immunoglobulin G (1:5,000; BS12478; Bio-World) for 2 h at room temperature. Following washing, protein bands were detected with chemiluminescent HRP substrate (SuperSignal West Pico; Thermo Fisher Scientific, Inc.) for 5 min at room temperature in the dark and exposed to X-ray film (Fujifilm Corp, Tokyo, Japan). The band intensities were analyzed using Quantity One software 4.6.2 (Bio-Rad Laboratories, Inc.) and normalized to the β-actin loading control.

TUNEL assay

For the TUNEL assay, a cell death detection kit was used (Promega Corp, Madison, WI, USA). Briefly, the sections were deparaffinized with xylene and rehydrated according to standard protocols, and were then incubated with proteinase-K for 20 min at room temperature followed by 3 washes in PBS. Subsequently, the sections were covered with equilibration buffer for 10 min at room temperature, followed by incubation with the rTdT incubation buffer (50 µl) in the dark for 1 h at 37°C. The slides were then immersed in 2X SSC buffer for 15 min and washed with PBS. Sections were counterstained with 4′,6′-diamino-2-phenylindole (DAPI; 1:1,000; Sigma-Aldrich, St. Louis, MO, USA), washed in PBS and mounted. Fluorescence microscopy was conducted using an Olympus BX51 microscope equipped with a mercury lamp. TUNEL-positive cells were normalized to the DAPI stained cells. The immunoreactive cells from 9 random fields (3 fields per sample, 3 samples per group) were counted using a 20x objective by an observer blinded to the treatment groups.

Statistical analysis

The data were presented as the mean ± standard deviations and analyzed using SPSS software, version 16.0 (SPSS, Inc., Chicago, IL, USA). A two-way analysis of variance was used to determine the significant differences between strain and age followed by a Tukey-Kramer post-hoc test to evaluate the statistical significance between the hypertensive state and age groups. P<0.05 was considered to indicate a statistically significant difference.

Results

Blood pressure

The SBP levels in the 16, 32 and 64-week-old SHR groups averaged 127.1±6.9, 184.4±8.1 and 193.9±11.3 mmHg, respectively. In aged-matched WKY rats, the SBP levels were significantly lower (102.9±4.7, 103.5±3.6 and 110.0±9.8 mmHg, respectively; P<0.01). A progressive augmentation of SBP was also observed in the SHR group with increasing age. Compared with the 16-week-old group, the 32 and 64-week-old groups showed significantly upregulated SBP levels (P<0.01). However, there was no significant difference between the two groups. Systolic pressure values were similar in the WKY groups of different ages.

Quantitative analysis of activated astrocytes

Activated astrocytes are markers of brain damage. The sections processed for GFAP immunohistochemistry exhibited thin, shallow, dark-brown astrocytes with slender branches in the WKY group (Fig. 1D). The SHR group exhibited increased occurrence of activated astrocytes, with large cell bodies and thick branches (Fig. 1C). Although the number of astrocytes was similar in the SHR and WKY groups of different ages, the ratio of activated astrocytes was increased progressively with age in the CA1 subfield in the SHR and WKY groups. The WKY group had 8.8, 12.9 and 21.8% activated astrocytes in the 16, 32 and 64-week-old groups, respectively (Fig. 1D–F). Compared with the age-matched WKY group, the SHRs exhibited a significant increase in the proportion of activated astrocytes (9.1, 25.3 and 53.1%, respectively; Fig. 1A–C). Apart from the 16-week-old group, the 32 and 64-week-old SHR groups were significantly different compared with the age-matched WKY groups (P<0.01). In addition, there were significant differences between the three SHR groups (P<0.01). The parameters investigated were similar in the 16 and 32-week-old WKY rats. The 64-week-old WKY group was significantly different compared with the 16 and 32-week-old WKY rats (P<0.05; Fig. 1G).

Quantitative analysis of TUNEL-positive cells

Sections processed for TUNEL staining revealed a few profiles of TUNEL-positive nuclei in the hippocampus of the 16 week SHR and WKY rats (13.7 and 13.4%, respectively; Fig. 2A and D). No statistically significant alteration was observed between the 16 week SHR and WKY rats. The apoptotic ratio of cells was reduced in the SHR group in comparison with the age-matched WKY rats. In the hippocampus of the 32-week-old SHRs, 24.9% of nuclei were TUNEL-positive (Fig. 2B). This number was reduced to 14.5% in the 32-week-old WKY group (Fig. 2E). A total of 46.8% TUNEL-positive nuclei were observed in the CA1 subfield of the 64-week-old SHRs (Fig. 2C), while the 64-week-old WKY group exhibited 22.2% TUNEL-positive nuclei (Fig. 2F). Apart from the 16-week-old group, the 32 and 64-week-old SHR groups were significantly different compared with the age-matched WKY groups (P<0.01). In addition, there were significant differences between the three SHR groups (P<0.01; Fig. 2G).

RT-qPCR assay

The expression of PPAR-γ mRNA in the SHR group reduced progressively with increasing age (P<0.01, compared within SHR groups). Compared with the age-matched WKY group, the 32 and 64-week-old SHR groups demonstrated a significantly reduced level of PPAR-γ mRNA expression (P<0.01). In addition, aging resulted in an age-dependent reduction of PPAR-γ mRNA expression levels in the WKY group. The 64-week-old WKY group were observed to be significantly different compared with the 16-week-old WKY group (P<0.05; Fig. 3A).

The expression levels of Bax and caspase-3 mRNA increased progressively with increasing age (P<0.01, compared within SHR groups). The 32 and 64-week-old SHR groups were significantly different compared with the age-matched WKY groups (P<0.01; Fig. 3B and C). Compared with the 16-week-old group, the 64-week-old group exhibited a significant difference in caspase-3 mRNA expression levels (P<0.01; Fig. 3B). Aging resulted in an age-dependent reduction of Bcl-2 mRNA expression levels in the SHR group (P<0.01, 16 vs. 32 and 64 weeks). Compared with the age-matched WKY groups, the Bcl-2 mRNA expression levels were observed to be significantly different in the 32 and 64-week-old SHR groups (P<0.01; Fig. 3D).

Western blot analysis

As shown in Fig. 4, the protein expression levels of PPAR-γ were progressively reduced with increasing age in the SHR groups (P<0.01, 16 vs. 32 week; P<0.05, 32 vs. 64 week). Apart from the 16-week-old group, the 32 and 64-week-old SHR groups showed a significant difference compared with the age-matched WKY groups (P<0.01). In addition, aging resulted in an age-dependent reduction of PPAR-γ protein expression levels in the WKY groups. However, only the 64-week-old WKY group was significantly different compared with the 16-week-old WKY group (P<0.05).

iNOS and gp47phox are markers used for assessing oxidative stress in brain tissues. As shown in Fig. 5, aging resulted in a significant upregulation of the protein expression levels of iNOS (P<0.01, 16 vs. 32 week; P<0.05, 32 vs. 64 week) and gp47phox (P<0.01, compared within SHR groups) in the SHR groups. Compared with the age-matched WKY group, the 32 and 64-week-old SHR groups exhibited significant differences in iNOS (P<0.05 and P<0.01, respectively) and gp47phox (P<0.01) expression levels. In addition, the 64-week-old WKY group showed significantly increased expression of gp47phox and iNOS protein compared with the 16-week-old WKY group (P<0.05).

The expression of Bax, Bcl-2 and caspase-3

As shown in Fig. 6, the protein expression levels of Bax, caspase-3 and the ratio of Bax/Bcl-2 were progressively increased with increasing age in SHR group (P<0.01). Apart from the 16-week-old group, the levels in the 32 and 64-week-old SHR groups were all significantly different compared with the age-matched WKY groups (P<0.01). In addition, the expression levels of Bax and caspase-3 protein in the 64-week-old WKY group were significantly increased compared with the 16-week-old WKY group (P<0.05, P<0.01, respectively). However, the expression of Bcl-2 protein was progressively reduced with increasing age in the SHR groups (P<0.05, 16 vs. 32 week; P<0.01, 16 vs. 64 week), but there was not a significant difference between the 32 and 64-week-old SHR groups. Compared with the age-matched WKY group, the 32 and 64-week-old SHR groups exhibited significantly reduced Bcl-2 protein expression (P<0.05, P<0.01, respectively).

Discussion

The main findings of the current study was that SHRs exhibited an age-dependent increase in apoptotic cells in the hippocampus and that this was accompanied by increased levels of oxidative stress and reduced expression of PPAR-γ. In addition, the degree of hypertensive brain damage, as indicated by activated astrocytes, was increased with increasing age, and the abnormally elevated Bax/Bcl-2 ratio may be involved in this process. To the best of our knowledge, the majority of previous studies assessing age-related oxidative stress levels and damage have been conducted in young/adult SHR models, ranging from 6–26 weeks of age (3,10,11). Few studies have examined the degree of age-related brain damage in the hippocampi of 32 and 64-week-old SHRs. The current study indicates that both aging and hypertension are able to exacerbate brain damage in SHRs, which was shown by the increased proportion of apoptotic cells and the expression of oxidative stress markers. The reduced expression of PPAR-γ may contribute to the age-related brain damage observed in SHRs.

SHRs represent a model of chronic hypertension, sharing several similarities with human hypertension and exhibiting behavioral and brain morphological alterations when compared with age-matched normotensive animals (16). Previous studies have observed significant reductions in the volume of grey matter, and a reduction in the number of neurons in the CA1 subfield of hippocampi in 6–7 month old SHRs, indicating that the volume reduction in grey matter potentially reflects neuronal loss occurring in certain hippocampal regions in SHRs (17,18). These previous studies suggest that depressed neuronal activity in SHRs occurs around 6 months of age (6,19). In the present study, no significant difference was observed in the ratio of TUNEL-positive cells between 16-week-old SHR and WKY groups, while the 32 and 64-week-old SHR groups showed a significant increase in the proportion of TUNEL-positive cells in the CA1 subfield of the hippocampus. The number of neurons in a given cerebral area is associated with the capability of the nervous system to receive, analyze and store information (20). The observation of increased levels of apoptotic cells in the hippocampus of SHR and aged WKY rats suggests that brain function may undergo a progressive impairment.

The specificity of the TUNEL technique for apoptosis has been questioned, as it may detect both apoptotic and necrotic nuclei (21). Although based on the TUNEL data it is not possible to conclude whether the CA1 neurons were undergoing apoptosis and/or necrosis, the demonstration of TUNEL-positive nuclei in the current study suggests the occurrence of cell death in the hippocampi of SHRs (22). This observation is consistent with a previous report which showed that excessive reactive oxygen species (ROS) induced by hypertension damaged neuronal components and lead to neuronal loss via apoptosis or necrosis (23). The demonstration of GFAP-immunoreactive astrocytes in the present study strengthens the suggestion of neuronal damage in the hippocampus of SHRs. GFAP is a cell-specific marker which distinguishes astrocytes from other glial cells. Increased expression of GFAP is considered to be an indicator of brain injury (24). The increased number of activated astrocytes in 32 and 64-week-old SHRs potentially indicates the detrimental effects of hypertension on the brain, and the attempt of astrocytes to protect the neuronal microenvironment.

Oxidative stress is suggested to be an important factor in aging and numerous age-related diseases (25), including essential hypertension (10). This is based on the presence of increased production of ROS, reduced NO synthesis and the reduced bioavailability of antioxidants, which have been verified in experimental and human hypertension (10,26). To date, oxidative stress has been demonstrated in patients with hypertension (27,28) and experimental hypertensive models (2932). As the most active organ with the greatest oxygen consumption and richness of fatty acids, the brain is more susceptible to oxidative stress injury than other organs, thus suggesting that oxidative stress may serve an important role in hypertension-associated brain damage (33). This is supported by evidence that the activation of cerebral nicotinamide adenine dinucleotide phosphate (NADPH) oxidase preceded cerebral inflammation and cellular apoptosis (34). gp47phox is an important subunit of the NADPH oxidase complex, and functions in the spatial organization of the various subunits of the enzyme (11,35). It has been observed that the enhancement of NADPH oxidase activity in SHRs was associated with an increase in gp47phox expression in vessels (11), which indicates that gp47phox is an important marker for assessing oxidative stress injury in SHRs. In addition, iNOS has been considered to be a marker of oxidative stress due to its ability to generate toxic levels of nitric oxide, which subsequently lead to cellular apoptosis or necrosis in the brain (36,37). In the current study, the increase in the protein expression levels of gp47phox and iNOS provides additional support for increased ROS production and oxidative stress in the brains of SHRs and aged WKY rats.

PPAR-γ is a ligand-activated transcription factor and serves beneficial roles in inhibiting inflammation and tissue injury (38). Recently, PPAR-γ has been suggested to be involved in the oxidative stress response. Oxidative stress can attenuate PPAR-γ expression and activity through the suppression of PPAR-γ transcription (13). A previous study reported that once activated, the PPAR-γ agonist was able to protect QZG cells against oxidative stress injury (14). Therefore, PPAR-γ may participate in oxidative stress-related injury and exert beneficial effects. To the best of our knowledge, a systematic analysis of the expression profile of PPAR-γ protein in SHRs of different ages has not been conducted. Diep et al (39) reported that PPAR-γ expression increases with age during the development of hypertension. Wu et al (40) observed that the basal protein expression levels of PPAR-γ in vascular tissues did not differ between the prehypertensive stage (5 weeks) or the evolving hypertensive state (13 weeks) of SHRs compared with age-matched WKY rats, while the protein levels of PPAR-γ were lower in 21-week-old SHRs compared with age-matched WKY rats (40). In the present study, it was observed that PPAR-γ expression was reduced with increasing age in the hippocampi of SHRs. Whilst this conflicts with the previous reports that PPAR-γ expression is elevated in the brains of SHRs (41), the difference may due to the increased age of the SHRs used in the present study. The current study demonstrated that SHRs exhibit an age-dependent reduction in PPAR-γ expression in the hippocampus, which may be involved in the hypertension-associated oxidative stress injury.

Bax is a member of the Bcl-2 family, and promotes apoptosis, while Bcl-2 blocks cell death (42). In a previous study, 32-week-old SHRs exhibited increased mRNA expression of Bax and reduced mRNA expression of Bcl-2 (43). The Bax/Bcl-2 ratio is a widely used parameter to determine cell susceptibility to apoptosis, and in the present study, an age-dependent increase in the Bax/Bcl-2 ratio was observed in the 32 and 64-week-old SHRs. The increase in oxidative stress markers and apoptosis regulatory proteins observed in the hippocampi of SHRs may be, at least in part, responsible for the increased proportion of apoptotic cells in these rats. However, the present study also observed that the 64-week-old WKY rats exhibited significantly increased active astrocytes and TUNEL-positive cells compared with 16 and 32-week-old WKY rats, which was accompanied by increased levels of oxidative stress markers and proapoptotic proteins. We hypothesized that the results reported in the present study may arise for two reasons: i) aging is an important factor in inducing cellular apoptosis; and ii) normotensive WKY rats may develop obesity with as they age. Common mechanisms leading to oxidative stress may underlie hypertension and obesity (44).

In conclusion, the present study demonstrated that SHRs and aged WKY rats exhibited increased apoptotic cells in the hippocampus, with this accompanied by increased levels of oxidative stress, suggesting the possible role of oxidative stress in hypertension-associated brain damage. Aging resulted in an age-dependent increase in oxidative stress and apoptotic cells in the hippocampi of SHRs. Reduced mRNA and protein expression of PPAR-γ the hippocampus may be involved in hypertension-associated oxidative stress injury. Further studies should be undertaken to further dissect the mechanisms and signaling pathways involved.

Acknowledgments

The authors would like to thank Professor Yi Zhu for his excellent technical assistance. The current study was supported by the National Natural Science Foundation of China (grant no. 81070219).

References

1 

Blackwood WS, Maudgal DP, Pickard RG, Lawrence D and Northfield TC: Cimetidine in duodenal ulcer. Controlled trial Lancet. 2:174–176. 1976.

2 

Kilander L, Nyman H, Boberg M, Hansson L and Lithell H: Hypertension is related to cognitive impairment: A 20-year follow-up of 999 men. Hypertension. 31:780–786. 1998. View Article : Google Scholar : PubMed/NCBI

3 

Sabbatini M, Tomassoni D and Amenta F: Hypertensive brain damage: Comparative evaluation of protective effect of treatment with dihydropyridine derivatives in spontaneously hypertensive rats. Mech Ageing Dev. 122:2085–2105. 2001. View Article : Google Scholar : PubMed/NCBI

4 

Harrison DG and Gongora MC: Oxidative stress and hypertension. Med Clin North Am. 93:621–635. 2009. View Article : Google Scholar : PubMed/NCBI

5 

Brown RC and Davis TP: Calcium modulation of adherens and tight junction function: A potential mechanism for blood-brain barrier disruption after stroke. Stroke. 33:1706–1711. 2002. View Article : Google Scholar : PubMed/NCBI

6 

Sabbatini M, Strocchi P, Vitaioli L and Amenta F: The hippocampus in spontaneously hypertensive rats: A quantitative microanatomical study. Neuroscience. 100:251–258. 2000. View Article : Google Scholar : PubMed/NCBI

7 

Geinisman Y, Detoledo-Morrell L, Morrell F and Heller RE: Hippocampal markers of age-related memory dysfunction: Behavioral, electrophysiological and morphological perspectives. Prog Neurobiol. 45:223–252. 1995. View Article : Google Scholar : PubMed/NCBI

8 

Schmidt-Kastner R, Ophoff BG and Hossmann KA: Pattern of neuronal vulnerability in the cat hippocampus after one hour of global cerebral ischemia. Acta Neuropathol. 79:444–455. 1990. View Article : Google Scholar : PubMed/NCBI

9 

Barone FC, Price WJ, White RF, Willette RN and Feuerstein GZ: Genetic hypertension and increased susceptibility to cerebral ischemia. Neurosci Biobehav Rev. 16:219–233. 1992. View Article : Google Scholar : PubMed/NCBI

10 

Touyz RM: Reactive oxygen species, vascular oxidative stress and redox signaling in hypertension: What is the clinical significance? Hypertension. 44:248–252. 2004. View Article : Google Scholar : PubMed/NCBI

11 

Amoureux S, Lorgis L, Sicard P, Girard C, Rochette L and Vergerly C: Vascular BDNF expression and oxidative stress during aging and the development of chronic hypertension. Fundam Clin Pharmacol. 26:227–234. 2012. View Article : Google Scholar

12 

Fuenzalida K, Quintanilla R, Ramos P, Piderit D, Fuentealba RA, Martinez G, Inestrosa NC and Bronfman M: Peroxisome proliferator-activated receptor gamma up-regulates the Bcl-2 anti-apoptotic protein in neurons and induces mitochondrial stabilization and protection against oxidative stress and apoptosis. J Biol Chem. 282:37006–37015. 2007. View Article : Google Scholar : PubMed/NCBI

13 

Blanquicett C, Kang BY, Ritzenthaler JD, Jones DP and Hart CM: Oxidative stress modulates PPAR gamma in vascular endothelial cells. Free Radic Biol Med. 48:1618–1625. 2010. View Article : Google Scholar : PubMed/NCBI

14 

Li WL, Liang X, Wang X, Zhang XD, Liu R, Zhang W, Chen HL, Qin XJ, Bai H and Hai CX: Protective effect of the peroxisome proliferator-activated receptor (PPAR)-γ, ligand rosiglitazone on tert-butyl hydroperoxide-induced QZG cell injury. Exp Toxicol Pathol. 63:527–533. 2011. View Article : Google Scholar

15 

Fleige S, Walf V, Huch S, Prgomet C, Sehm J and Pfaffl MW: Comparison of relative mRNA quantification models and the impact of RNA integrity in quantitative real-time RT-PCR. Biotechnol Lett. 28:1601–1613. 2006. View Article : Google Scholar : PubMed/NCBI

16 

Lehr RP Jr, Browning RA and Myers JH: Gross morphological brain differences between Wistar-Kyoto and spontaneously hypertensive rats. Clin Exp Hypertens. 2:123–127. 1980. View Article : Google Scholar : PubMed/NCBI

17 

Tajima A, Hans FJ, Livingstone D, Wei L, Finnegan W, Demaro J and Fenstermacher J: Smaller local brain volumes and cerebral atrophy in spontaneously hypertensive rats. Hypertension. 21:105–111. 1993. View Article : Google Scholar : PubMed/NCBI

18 

Sabbatini M, Catalani Å, Consoli C, Marletta N, Tomassoni D and Avola R: The hippocampus in spontaneously hypertensive rats: An animal model of vascular dementia? Mech Ageing Dev. 123:547–559. 2002. View Article : Google Scholar : PubMed/NCBI

19 

Tomassoni D, Avola R, Mignini F, Parnetti L and Amenta F: Effect of treatment with choline alphoscerate on hippocampus microanatomy and glial reaction in spontaneously hypertensive rats. Brain Res. 1120:183–190. 2006. View Article : Google Scholar : PubMed/NCBI

20 

Napoleone P, Ferrante F, Ghirardi O, Ramacci MT and Amenta F: Age-dependent nerve cell loss in the brain of Sprague-Dawley rats: Effect of long term acetyl-L-carnitine treatment. Arch Gerontol Geriatr. 10:173–185. 1990. View Article : Google Scholar : PubMed/NCBI

21 

Charriaut-Marlangue C and Ben-Ari Y: A cautionary note on the use of the TUNEL stain to determine apoptosis. Neuroreport. 7:61–64. 1995. View Article : Google Scholar : PubMed/NCBI

22 

Hamet P, Richard L, Dam TV, Teiger E, Orlov SN, Gaboury L, Gossard F and Tremblay J: Apoptosis in target organs of hypertension. Hypertension. 26:642–648. 1995. View Article : Google Scholar : PubMed/NCBI

23 

Lo EH, Dalkara T and Moskowitz MA: Mechanisms, challenges and opportunities in stroke. Nat Rev Neurosci. 4:399–415. 2003. View Article : Google Scholar : PubMed/NCBI

24 

Takamiya Y, Kohsaka S, Toya S, Otani M and Tsukada Y: Immunohistochemical studies on the proliferation of reactive astrocytes and the expression of cytoskeletal proteins following brain injury in rats. Brain Res. 466:201–210. 1988. View Article : Google Scholar : PubMed/NCBI

25 

Kregel KC and Zhang HJ: An integrated view of oxidative stress in aging: Basic mechanisms, functional effects and pathological considerations. Am J Physiol Regul Integr Comp Physiol. 292:R18–R36. 2007. View Article : Google Scholar

26 

Tanito M, Nakamura H, Kwon YW, Teratani A, Masutani H, Shioji K, Kishimoto C, Ohira A, Horie R and Yodoi J: Enhanced oxidative stress and impaired thioredoxin expression in spontaneously hypertensive rats. Antioxid Redox Signal. 6:89–97. 2004. View Article : Google Scholar : PubMed/NCBI

27 

Higashi Y, Sasaki S, Nakagawa K, Matsuura H, Oshima T and Chayama K: Endothelial function and oxidative stress in renovascular hypertension. N Engl J Med. 346:1954–1962. 2002. View Article : Google Scholar : PubMed/NCBI

28 

Zhou L, Xiang W, Potts J, Floyd M, Sharan C, Yang H, Ross J, Nyanda AM and Guo Z: Reduction in extracellular superoxide dismutase activity in African-American patients with hypertension. Free Radic Biol Med. 41:1384–1391. 2006. View Article : Google Scholar : PubMed/NCBI

29 

Dobrian AD, Schriver SD, Khraibi AA and Prewitt RL: Pioglitazone prevents hypertension and reduces oxidative stress in diet-induced obesity. Hypertension. 43:48–56. 2004. View Article : Google Scholar

30 

Nishiyama A, Yao L, Nagai Y, Miyata K, Yoshizumi M, Kagami S, Kondo S, Kiyomoto H, Shokoji T, Kimura S, et al: Possible contributions of reactive oxygen species and mitogen-activated protein kinase to renal injury in aldosterone/salt-induced hypertensive rats. Hypertension. 43:841–848. 2004. View Article : Google Scholar : PubMed/NCBI

31 

Pechanova O, Zicha J, Kojsova S, Dobesová Z, Jendeková L and Kunes J: Effect of chronic N-acetylcysteine treatment on the development of spontaneous hypertension. Clin Sci (Lond). 110:235–242. 2006. View Article : Google Scholar

32 

Kushiro T, Fujita H, Hisaki R, Asai T, Ichiyama I, Kitahara Y, Koike M, Sugiura H, Saito F, Otsuka Y and Kanmatsuse K: Oxidative stress in the Dahl salt-sensitive hypertensive rat. Clin Exp Hypertens. 27:9–15. 2005. View Article : Google Scholar : PubMed/NCBI

33 

Poulet R, Gentile MT, Vecchione C, Distaso M, Aretini A, Fratta L, Russo G, Echart C, Maffei A, De Simoni MG and Lembo G: Acute hypertension induces oxidative stress in brain tissues. J Cereb Blood Flow Metab. 26:253–262. 2006. View Article : Google Scholar

34 

Yamamoto E, Tamamaki N, Nakamura T, Kataoka K, Tokutomi Y, Dong YF, Fukuda M, Matsuba S, Ogawa H and Kim-Mitsuyama S: Excess salt causes cerebral cellular apoptosis and inflammation in stroke-prone hypertensive rats through angiotensin II-induced NADPH oxidase activation. Stroke. 39:3049–3056. 2008. View Article : Google Scholar : PubMed/NCBI

35 

Infanger DW, Sharma RV and Davisson RL: NADPH oxidases of the brain: Distribution, regulation and function. Antioxid Redox Signal. 8:1583–1596. 2006. View Article : Google Scholar : PubMed/NCBI

36 

Jiang T, Gao L, Shi J, Lu J, Wang Y and Zhang Y: Angiotensin-(1-7) modulates renin-angiotensin system associated with reducing oxidative stress and attenuating cellular apoptosis in the brain of hypertensive rats. Pharmacol Res. 67:84–93. 2013. View Article : Google Scholar

37 

Yamagata K, Tagami M and Yamori Y: Neuronal vulnerability of stroke-prone spontaneously hypertensive rats to ischemia and its prevention with antioxidants such as vitamin E. Neuroscience. 170:1–7. 2010. View Article : Google Scholar : PubMed/NCBI

38 

Narala VR, Subramani PA, Narasimha VR, Shaik FB and Panati K: The role of nitrated fatty acids and peroxisome proliferator-activated receptor gamma in modulating inflammation. Int Immunopharmacol. 23:283–287. 2014. View Article : Google Scholar : PubMed/NCBI

39 

Diep QN and Schiffrin EL: Increased expression of peroxisome proliferator-activated receptor-alpha and -gamma in blood vessels of spontaneously hypertensive rats. Hypertension. 38:249–254. 2001. View Article : Google Scholar : PubMed/NCBI

40 

Wu L, Wang R, De Champlain J and Wilson TW: Beneficial and deleterious effects of rosiglitazone on hypertension development in spontaneously hypertensive rats. Am J Hypertens. 17:749–756. 2004. View Article : Google Scholar : PubMed/NCBI

41 

Sun L, Gao YH, Tian DK, Zheng JP, Zhu CY, Ke Y and Bian K: Inflammation of different tissues in spontaneously hypertensive rats. Sheng Li Xue Bao. 58:318–323. 2006.PubMed/NCBI

42 

Youle RJ and Strasser A: The BCL-2 protein family: Opposing activities that mediate cell death. Nat Rev Mol Cell Biol. 9:47–59. 2008. View Article : Google Scholar

43 

Li Y, Duan Z, Gao D, Huang S, Yuan H and Niu X: The new role of LOX-1 in hypertension induced cellular apoptosis. Biochem Biophys Res Commun. 425:735–740. 2012. View Article : Google Scholar : PubMed/NCBI

44 

Simao S, Gomes P, Pinto V, Silva E, Amaral JS, Igreja B, Afonso J, Serrão MP, Pinho MJ and Soares-da-Silva P: Age-related changes in renal expression of oxidant and antioxidant enzymes and oxidative stress markers in male SHR and WKY rats. Exp Gerontol. 46:468–474. 2011. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

March-2016
Volume 13 Issue 3

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
Li Y, Liu J, Gao D, Wei J, Yuan H, Niu X and Zhang Q: Age‑related changes in hypertensive brain damage in the hippocampi of spontaneously hypertensive rats. Mol Med Rep 13: 2552-2560, 2016
APA
Li, Y., Liu, J., Gao, D., Wei, J., Yuan, H., Niu, X., & Zhang, Q. (2016). Age‑related changes in hypertensive brain damage in the hippocampi of spontaneously hypertensive rats. Molecular Medicine Reports, 13, 2552-2560. https://doi.org/10.3892/mmr.2016.4853
MLA
Li, Y., Liu, J., Gao, D., Wei, J., Yuan, H., Niu, X., Zhang, Q."Age‑related changes in hypertensive brain damage in the hippocampi of spontaneously hypertensive rats". Molecular Medicine Reports 13.3 (2016): 2552-2560.
Chicago
Li, Y., Liu, J., Gao, D., Wei, J., Yuan, H., Niu, X., Zhang, Q."Age‑related changes in hypertensive brain damage in the hippocampi of spontaneously hypertensive rats". Molecular Medicine Reports 13, no. 3 (2016): 2552-2560. https://doi.org/10.3892/mmr.2016.4853