Open Access

Dynamic effects of chronic unpredictable mild stress on the hippocampal transcriptome in rats

  • Authors:
    • Feng Li
    • Ying Wang
    • Xue Wang
    • Yun Zhao
    • Fang Xie
    • Ling-Jia Qian
  • View Affiliations

  • Published online on: February 1, 2022     https://doi.org/10.3892/mmr.2022.12626
  • Article Number: 110
  • 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

Stress causes extensive changes in hippocampal genomic expression, leading to changes in hippocampal structure and function. The dynamic changes in hippocampal gene expression caused by stress of different durations are still unknown. mRNA sequencing was used to analyze the hippocampal transcriptome of rats subjected to chronic unpredictable mild stress (CUMS) of different durations. Compared with the control, 501, 442 and 235 differentially expressed genes (DEGs) were detected in the hippocampus of rats subjected to CUMS for 3 days and 2 and 6 weeks, respectively. Gene Ontology (GO) analysis was used to determine the potential mechanism underlying the dynamic harmful effects of stress on the hippocampus; Certain GO terms of the down‑regulated DEGs in CUMS (3 days) rats were also found in the up‑regulated DEGs in CUMS (6 weeks) rats. These results showed opposing regulation patterns of DEGs between CUMS at 3 days and 6 weeks, which suggested a functional change from adaptation to damage in during the early and late stages of chronic stress. GO analysis for upregulated genes in rats subjected to CUMS for 3 days and 2 weeks suggested significant changes in ‘extracellular matrix’ and ‘wound healing’. Upregulated genes in rats subjected to CUMS for 2 weeks were involved in changes associated with visual function. GO analysis of DEGs in rats subjected to CUMS for 6 weeks revealed increased expression of genes associated with ‘apoptotic process’ and ‘aging’ and decreased expression of those associated with inhibition of cell proliferation and cell structure. These results suggest that the early and middle stages of chronic stress primarily promote adaptive regulation and damage repair in the organism, while the late stage of chronic stress leads to damage in the hippocampus.

Introduction

Chronic stress is a non-specific complex psychosomatic process that is a key risk factor for depression (1), anxiety (2), drug abuse (3), cardiovascular disease (4), ulcers (5) and cancer (6). Therefore, investigating the positive or negative responses to stressors is important to reveal the pathogenesis of stress-associated disease. The body's response to stressors depends on the nature, intensity and duration of the stressor, forming an ‘inverted U’-shaped continuous quantitative effect curve (7). In long-term stress, early stress causes the body to produce compensatory responses, such as increased alertness, sensitivity; as duration of stress increases, the body systems act in concert to maintain homeostasis, thus actively adapting to the stressor. The cumulative effect of further stress leads to a disturbance of homeostasis in the body, resulting in pathophysiological damage (1).

The hippocampus is responsible for learning, memory, cognition and emotion; it is also one of the key brain areas that mediate stress reactions (8). Prolonged stress results in decreased hippocampal nerve cell plasticity, disequilibrium between hippocampal apoptosis and regeneration, leading to nerve cell atrophy and loss and causing structural and functional damage in the hippocampus (9,10). Studies have investigated the effects of different stress patterns on gene expression in the hippocampus (11,12), but, to the best of our knowledge, there are no reports on dynamic changes in gene expression in the hippocampus during chronic stress.

To determine the effects of stress on hippocampal structure, the present study investigated transcriptomic changes in the rat hippocampus during chronic unpredictable mild stress (CUMS) using mRNA sequencing (seq) technology. The CUMS rodent model uses both physiologically and psychologically stressful stimulations to mimic adverse stress from negative life events and induces decreased locomotion and anhedonia, similar to symptoms of depression in humans, which makes it a widely used animal model to study the molecular basis of stress in the brain (13). The present study investigated gene expression in the hippocampus of rats in the early, middle and late stages of CUMS to clarify continuous gene expression changes in the hippocampus in CUMS and reveal the pathophysiological mechanism of stress injury.

Materials and methods

Animals

A total of 36 male Sprague Dawley (SD) rats (age, 2 months; weight, 180–220 g) were obtained from SBF (Beijing) Biotechnology Co., Ltd. All animals were housed in groups of three or four in plastic cages and maintained under standard laboratory conditions (12/12-h light/dark cycle, 22±2°C, relative humidity 40–70%, ad libitum access to food and water).

CUMS procedure

In brief, SD male rats were randomly divided into four groups (n=9 per group) as follows: i) Control and CUMS for ii) 3 days and iii) 2and iv) 6 weeks (Fig. 1). The CUMS model was established as previously reported (14): i) Overnight light exposure, ii) darkness for 12 h, iii) wet bedding for 12 h, iv) removal of bedding for 12 h, v) deprivation of food for 12 h, vi) deprivation of water for 12 h, vii) soaking for 3 min in cold water (4°C), viii) physical restraint for 4 h, ix) cage vibration (70 rpm) for 2 h and x) electrical shock to the feet for 5 min (0.2 mA). Stressors were randomly scheduled to produce an unexpected mild stress effect. A combination of two different stressors was applied separately each day and the same stressor was not repeated within 3 days. Control animals were left uninterrupted except for regular cage cleaning. The effectiveness of the applied stress procedure for induction of the stress response was confirmed by cognitive behavior tests.

Open field test

The test was performed in a black wooden box measuring 100×100×45 cm. The floor was divided into 25 equal squares (20×20 cm). At the beginning of the experiment, rats were placed individually in the center of the open field and ambulation (number of squares crossed) and head-rise frequency were observed for 5 min using an ANY-maze behavior analysis system (Stoelting Co.). To ensure that the test results were free from any previous residual effect, the inside of the box was cleaned before each test. An open field test score was calculated as the sum of ambulation and immobility frequency.

Novel object recognition test

The experimental device was a black wooden box (60×60×45 cm). The test comprised three steps: Adaptation, training and testing. On the day before the experiment, rats were allowed to move freely in the experimental setup (without objects) for 10 min. During the training trial, two identical objects were placed on the left and right ends of a sidewall, and the rats were placed in the field with their back facing the two objects. The animal was placed in the center of the open field and allowed to explore freely for 10 min. During the test, one of the two objects was replaced with another object of different color and shape and the rat was returned to the same box for 10 min. The ANY-maze behavior analysis system recorded the time rats spent exploring the new and old object, denoted as Tnew and Told, respectively. The cognitive index (CI) was expressed as follows: CI=(Tnew-Told)/(Tnew + Told).

Morris water maze

The experimental device was a black circular plastic tank (diameter, 160 cm; depth, 75 cm) containing water at 22±2°C. The curtains around the tank were labeled with stickers of different shapes and colors serving as spatial reference cues. Rats were first required to complete 5 half-days of training at a frequency of once a day for 3 min each. During training, rats were placed into the maze in one of four quadrants facing the curtain and were trained to find a hidden platform submerged 2 cm beneath the surface of the water. If rats could not find the platform within 3 min, they were guided to the platform and kept there for 10 sec. During the test phase, the platform was removed and rats were placed into the water on the opposite quadrant. Each rat was allowed to swim in the water for 3 min, during which, the first time the rats found the original platform location and the number of platform passes within 3 min were recorded.

Sample preparation

Following behavioral observation, rats were decapitated following anesthesia with sodium pentobarbital (1%, 50 mg/kg, intraperitoneal). The whole brain was removed and washed with cold 0.9% saline. The hippocampus was dissected on an ice plate on a super-clean bench. The dissected hippocampi were placed into freezing EP tubes, frozen in liquid nitrogen and transferred into a −80°C low-temperature refrigerator for storage and use. RNA was extracted from hippocampal tissue using TRI reagent (Sigma-Aldrich; Merck KGaA) according to the manufacturer's protocol. Three hippocampal tissue samples were pooled as a single sample for the total transcript array analysis to increase the quantity of extracted RNA. RNA purity and quantification were evaluated using the NanoDrop 2000 spectrophotometer (Thermo Fisher Scientific, Inc.). RNA integrity was assessed using the Agilent 2100 Bioanalyzer (Agilent Technologies, Inc.). All samples selected for further analysis were of high quality (RNA integrity number >9, 260/280-2.1). Then the libraries were constructed using VAHTS Universal V6 RNA-seq Library Prep Kit for Illumina (catalog no. NR604-02; Nanjing Vazyme Biotech Co.) according to the manufacturer's instructions. The concentration of the library used for sequencing was ≥3.75 fmol/µl and was measured using a Qubit ssDNA Assay Kit (catalog no. Q10212) and Qubit Fluorometer (catalog no. Q33238) (both; Invitrogen; Thermo Fisher Scientific, Inc.). Transcriptome sequencing and analysis were performed by Shanghai Oe Biotech Co., Ltd.

RNA-seq and differentially expressed gene (DEG) analysis

The libraries were sequenced on MGI DNBSEQ-T7 platform using DNBSEQ-T7RS high-throughput sequencing kit (1000016102, MGI), and 150 bp paired-end reads were generated. A total of ~49.5 million raw reads were generated for each sample. Raw data (raw reads) of FASTQ format were processed using Trimmomatic (version 0.36) (15) and low-quality reads (adaptor and data of Q30 <85%) were removed to obtain the clean reads. Then ~49 million clean reads were retained for each sample for subsequent analysis.

The clean reads were mapped to the rat genome (mRatBN7.2) using Hierarchical Indexing for Spliced Alignment of Transcripts 2 (version 2.2.1.0) (16). Fragments per kilobase of transcript per million mapped reads (FPKM) of each gene was calculated using Cufflinks (version 2.2.1) (17) and the read count of each gene was obtained by HTSeqcount (version 0.9.1) (18). Differential expression analysis was performed using DESeq (2012) R package (19). P<0.05 and fold-change >2 or <0.5 was set as the threshold for significantly differential expression.

To identify the functional classes associated with DEGs, the online analysis tool Database for Annotation, Visualization, and Integrated Discovery 6.8 (david.ncifcrf.gov/summary.jsp) and Gene Ontology (GO) database were used to classify the biological process, molecular function, and cellular components of DEGs. GO annotations were obtained from GO database version 2.2 (20). Hypergeometric distributions were used to detect over- or under-represented biological process terms in the studied set compared with the population set. Terms with P<0.05 were considered to be significant.

Confirmation of gene expression levels using reverse transcription-quantitative (RT-q)PCR

All selected hippocampal RNA specimens were chosen from the animals used in the mRNA-seq study and were analyzed individually. Meeting three strict conditions [P<0.05, log2(fold-change)>1 and 2<average FPKM<50], seven genes were selected to verify the accuracy and reliability of mRNA-seq data. These genes were tested using SYBR-Green-based RT-qPCR in 96-well plates on a LightCycler 96 thermocycler (Roche Diagnostics GmbH). Primers (Table I) were designed using Primer-BLAST tool (National Center for Biotechnology Information) with the Rattus norvegicus RefSeq database (reference genome mRatBN7.2). The primers produced amplicons spanning exon-exon junctions and including all known alternatively spliced mRNA variants. A total of 1,000 ng total RNA from each sample was reverse-transcribed to cDNA according to the manufacturer's instructions (ABScript II RT Mix for qPCR, ABclonal Biotech Co., Ltd.). RT-qPCR was run using TB Green Premix Ex Taq II (Takara Bio, Inc.) according to the manufacturer's instructions. The volume of the PCR reaction system, cDNA sample, forward and reverse prime, and TB Green II were 10.0, 1.0, 0.2, 0.2 and 5.0 µl, respectively. The thermocycling conditions were as follows: Initial denaturation for 2 min at 95°C followed by 50 cycles of 95°C for 10 sec, 60°C for 10 sec and 72°C for 15 sec. All experiments were performed in triplicate. Raw cycle threshold values were calculated on LightCycler. Samples were analyzed by the ΔΔCq method. ΔCq values represent normalized target genes levels with respect to the internal control. Normalization was based on a single reference gene (β-actin). ΔΔCq values were calculated as the ΔCq of each test sample (CUMS) minus the mean ΔCq of the calibrator samples (ctrl) for each target gene. The fold change was calculated using the equation 2−ΔΔCq (21).

Table I.

Primer sequences for reverse transcription-quantitative PCR.

Table I.

Primer sequences for reverse transcription-quantitative PCR.

GeneForward sequence (5′-3′)Reverse sequence (5′-3′)Annealing temperature, °CProduct size, bp
Slc6a20CTA CAT CCT CAC GGG AAC GCGTG GCT GAC TTC GGT CTT TG60231
Bmp7CAT GGA CCC CAG AAC AAG CACTT TGG AGT CTT GGA GCG GT60123
PtgdsGTG CAG CCC AAC TTT CAA CACCG GAA CCA GCT TGA ATT GG6081
Hbb-b1TTG GCA GCC TCA GTG AAC TCCGAC AGA AGC TCT CTT GGG AAC A60248
Pdk4AAT GTG GTC CCT ACG ATG GCATC GCA GTT GGG GTC GAT AC60213
SccpdhCGA AAC CAG ATG AAC GGC ACTGC CCT TAT CGC CAA AAC CA60127
Npas4GTG GCA GCA CTA CCT GGA TTGTT TGT TGC CTG CAC TCT GG60265
ActbCTT CCT GGG TAT GGA ATC CTTCT TTA CGG ATG TCA ACG TC6080

[i] Slc6a20, solute carrier family 6 member 20; Bmp7, bone morphogenetic protein 7; Ptgds, prostaglandin D2 synthase; Hbb-b1, hemoglobin, β adult major chain; Sccpdh, saccharopine dehydrogenase (putative); Npas4, neuronal PAS domain protein 4, Actb, β-actin.

Statistical analysis

The results of behavioral tests were compared using Shapiro-Wilk test. The expression of each mRNA in the hippocampus of CUMS and control groups was compared using Mann-Whitney U test as data did not show a normal distribution. Nine independent experiments were carried out in each group in behavioral tests, while three independent experiments were conducted in each group in RT-qPCR experiments. All data are presented as the mean ± standard deviation. GraphPad Prism 8 software (ver. 8.4.2; GraphPad Software, Inc.) was used for statistical analysis and Spearman's correlation was used for the correlation analysis. P<0.05 was considered to indicate a statistically significant difference.

Results

Chronic stress induced time-dependent cognitive decline in rats

Three different behavioral tests were used to evaluate cognitive function in rats. The behavioral results showed that the open field score, object recognition CI and number of platform passes decreased with longer stress time, while the escape latency in the Morris water maze test increased gradually. The difference (P<0.05) first occurred in the open-field test at 2 weeks of CUMS. All behavioral performance associated with cognitive function decreased significantly (P<0.01) by week 6 of CUMS, suggesting that chronic stress induced cognitive dysfunction in rats (Fig. S1).

Number of DEGs decreases with the increase in stress time

Compared with the control, 501 DEGs were identified in CUMS (3 days) rats, of which 425 were up- and 76 were downregulated. A total of 442 DEGs in the CUMS (2 weeks) group were identified, of which 353 were up- and 89 were downregulated. A total of 235 DEGs in the CUMS (6 weeks) group were identified, of which 103 were up- and 132 were downregulated (Fig. 2; Table SI). As the duration of stress increased, the total number of DEGs decreased; this was accompanied by a decrease in up- and an increase in downregulated genes.

GO function analysis of DEGs suggests a functional change from adaptation to damage in the early and late stages of chronic stress

DEGs were divided into up- and downregulated groups for GO enrichment analysis. GO terms were divided into three categories: Biological processes, cellular components and molecular functions. The top 10 terms with the most significant P-values (<0.05) were selected from each category. Significant biological processes of upregulated DEGs of the hippocampus in CUMS (3 days) rats included ‘collagen fibril organization’, ‘oxygen transport’, ‘wound healing’, ‘cell-matrix adhesion’, ‘ammonium transmembrane transport’, ‘cerebrospinal fluid secretion’ and ‘blood vessel development’ (Fig. 3A), while those of downregulated DEGs involved ‘positive regulation of neuron projection development’, ‘inflammatory response’, ‘multicellular organism development’, ‘regulation of apoptotic process’, ‘protein phosphorylation’ and ‘oxidation-reduction process’ (Fig. 3B). Significant biological processes of upregulated DEGs in CUMS (2 weeks) rats included ‘visual perception’, ‘phototransduction’, ‘retina development in camera-type eye’, ‘oxygen transport’, ‘wound healing’, ‘regulation of cilium beat frequency’, ‘cerebrospinal fluid secretion’, ‘collagen fibril organization’, ‘detection of light stimuli involved in visual perception’ and ‘receptor-mediated endocytosis’ (Fig. 3C), while those of the downregulated DEGs in this group involved ‘protein phosphorylation’, ‘positive regulation of MAPK cascade’, ‘inflammatory response’, ‘apoptotic process’ and ‘intracellular protein transport’ (Fig. 3D). Certain biological processes were found after both 3 days and 2 weeks of CUMS. Significant biological processes of upregulated DEGs in CUMS (6 weeks) rats involved ‘translation’, ‘transmembrane transport’, ‘immune response’, ‘protein homooligomerization’, ‘regulation of apoptotic process’ and ‘aging’ (Fig. 3E), while those of the downregulated DEGs in this group included ‘positive regulation of blood pressure’, ‘receptor-mediated endocytosis’, ‘cell-cell adhesion’, ‘cell-cell signaling’, ‘proteolysis’, ‘angiogenesis’, ‘signal transduction’ and ‘oxidation-reduction process’ (Fig. 3F). Significant cellular components and molecular functions of DEGs are shown in the supplementary materials (Figs. S2 and S3). ‘Integral component of the plasma membrane’ was involved in the cellular component of downregulated DEGs in CUMS (3 days) rats and upregulated DEGs in CUMS (6 weeks) rats. ‘Protein binding’, ‘identical protein binding’, ‘metal ion binding’ and ‘ATP binding’ were involved in the molecular functions of downregulated DEGs in CUMS (3 days) rats and upregulated DEGs in CUMS (6 weeks) rats. Furthermore, significant cellular components of downregulated DEGs in CUMS (6 weeks) rats involved ‘synapse’, ‘dendrite’ and ‘axon’.

Results of RT-qPCR coincide with the changes in differential gene identified by mRNA-seq

A total of seven differentially expressed transcripts were verified by RT-qPCR. Slc6a20, Bmp7, Ptgds and Hbb-b1 were upregulated during CUMS for 3 days and 2 weeks, and gradually decreased to normal levels at 6 weeks of CUMS. These changes in differential gene detected in mRNA-seq data were confirmed by RT-qPCR. Similarly, Pdk4, Npas4 and Sccpdh were downregulated in the whole CUMS process and their changes identified by RT-qPCR were consistent with mRNA-seq data (Fig. 4A). A significant positive correlation between the expression level data from mRNA-seq and RT-qPCR was identified (r=0.891; P<0.001) and the slope of the correlation line was 0.915 (Fig. 4B).

Discussion

In the present study, gene expression profiling of the hippocampus in rats exposed to stress was investigated using mRNA-seq. The GO analysis of DEGs in CUMS for 3 days and 6 weeks shows an inversion of some biological functions. The downregulated genes in CUMS (3 days) rats shared some GO terms with upregulated genes in CUMS (6 weeks) rats, including ‘regulation of the apoptotic process’, ‘integral component of the plasma membrane’, ‘protein binding’, ‘identical protein binding’, ‘metal ion binding’ and ‘ATP binding’. These results suggested a functional change from adaptation to damage during the early and late stages of chronic stress. Other studies have also reported similar inverted changes in hormones and proteins during stress (2225). At the onset of stress, glucocorticoids (GCs) exert anti-apoptotic effects targeting protein Bcl-2, while prolonged stress increases free radicals and facilitates apoptosis (22). Moreover, as a key neurotrophic factor, brain-derived neurotrophic factor (BDNF) exhibits a transient increase in the hippocampus following acute stress, whereas chronic stress leads to downregulation of BDNF in the CA3 region of the hippocampus (2325). The mechanism underlying inverted changes during CUMS is still unknown. Differences between GC receptors [mineralocorticoid receptor (MR) and glucocorticoid receptor (GR)] may account for the inverted changes. GCs at low concentrations in the early stage of stress act via high-affinity MR, which will promote the increase of neuronal excitability. However, higher GC levels in the late stage activate low-affinity GR to exert different effects, such as inhabitation of proliferation and integration (26). Norepinephrine is secreted rapidly in the early stage of stress, while GC levels increase during long-term stress. In the hippocampal dentate gyrus, norepinephrine enhances excitability and synaptic plasticity, but the high levels of corticosterone suppress noradrenergic activity (27). The duration and antagonism of hormones during stress may contribute to the inversion of biological functions. Further studies are needed to investigate the mechanism underlying altered biological processes during chronic stress.

GO analysis of upregulated genes in CUMS (2 weeks) rats revealed vision-associated functional alterations, such as ‘visual perception’, ‘phototransduction’, ‘retina development in the camera-type eye’, ‘detection of light stimuli involved in visual perception’, ‘photoreceptor outer segment’, ‘photoreceptor inner segment’ and ‘retinal binding’. This indicated that visual function was significantly altered after 2 weeks of stress, when cognitive impairment had not yet developed. The association between changes in visual function and stress is still unclear. Certain studies have found that stress leads to attentional narrowing and enhanced ability for detection and observation (28,29). However, other studies have found decreased attention and accuracy of visual perceptual processing in stressful states (30,31). The effects of acute stress on visual function are controversial. Chronic stress is hypothesized to be the most common cause of depression. In addition to symptoms such as anxiety, depression is often accompanied by perceptual abnormality, including visual changes; such changes may comprise notable diminished visual acuity or increased sensitivity to light stimuli (32,33). Certain studies have reported visual impairment in patients with depression, including impaired ability to distinguish intensity of light in a room and the ability to read (32,34). Golomb et al (33) used a classical visual stimulus, the moving grating test, to test patients with major depressive disorder who had normal visual acuity and found that motion perception was enhanced in patients with depression compared with healthy individuals. The present findings suggested that visual function may be an early indicator of impairment due to stress. Further studies are needed to determine changes in visual perception during stress.

GO analysis of upregulated genes in CUMS (3 days and 2 weeks) rats showed functional changes in ‘extracellular matrix’ (ECM), ‘extracellular space’ and ‘extracellular region’. GO analysis of downregulated genes in these groups also showed functional changes in ‘extracellular space’ and ‘extracellular region’, suggesting an imbalance in synthesis and degradation of extracellular components of hippocampal tissue during the early and middle stages of CUMS. This imbalance might cause excessive deposition of ECM components in the hippocampal tissue (35), such as collagen, metalloproteinase, fibronectin, bone morphogenetic protein 7, nidogen 1 and elastin, which contributes to development of organ fibrosis (36). Collagen is a primary structural component of ECM and serves a key role in the cell cycle stimulated by mitogen (37). However, certain reports have shown that collagen inhibits cell proliferation (38,39). Excessive deposition of fibrous collagen, such as type I collagen, occurs during organ fibrosis, which leads to tissue and organ sclerosis and loss of function. Activation of type I collagen COL1A1 in the hippocampus contributes to the development of hippocampal sclerosis (40). The primary pathological manifestations of hippocampal sclerosis include hippocampal atrophy, neuronal loss, granule cell reorganization, alteration of interneuronal populations and chronic fibrillary gliosis centered on the pyramidal cell layer (41). GO analysis of upregulated DEGs in CUMS (3 days and 2 weeks) groups also indicated a wound healing function. Increased levels of collagen fibers serve an important role in tissue injury healing (42,43). Activation of ECM in the hippocampus during the early and middle stages of CUMS is hypothesized to contribute to self-repair of damaged hippocampal cells but also promotes hippocampal sclerosis, which leads to partial neuronal loss (35). ECM changes are often secondary to inflammatory reactions (36), which indicates that hippocampal inflammation occurs during the early and middle stages of CUMS. Excessive ECM deposition is a damage repair response to hippocampal cell damage and inflammation caused by stress, which contributes to the development of hippocampal sclerosis while promoting self-repair (44).

GO analysis of the upregulated DEGs in CUMS (6 weeks) rats involved changes in ‘aging’, suggesting that chronic stress shares certain pathways with aging or that long-term chronic stress promotes aging. Aging is often accompanied by pathological signs such as memory loss, Alzheimer's disease, dementia and depression, and chronic stress is one of the contributing factors to these pathological signs (45,46). The results of numerous animal stress experiments and clinical trials show that stress leads to changes in brain structure and function, cognitive decline and accelerated cellular aging (46,47). Sapolsky et al (48) proposed a GC cascade hypothesis of aging, suggesting that stress and GC accelerate the aging process in rats. The hippocampus is a target and a highly vulnerable area of GC in the brain. Elevated GC levels in the blood of elderly people leads to hippocampal atrophy, the degree of which is positively correlated with degree of GC elevation, resulting in hippocampal dysfunction and deficits in memory and learning and cognitive ability (49,50). In addition, one study found that chronic stress disrupts the normal processing of Aβ precursor protein, thus leading to the accumulation of Aβ in the brain (14). The development of physiological aging may interact with certain genetic background or negative environmental conditions (such as chronic stressors and stress in infancy and early childhood) to trigger dysregulation of the hypothalamus-pituitary-adrenal (HPA) axis (51) and changes in the neurotransmitter system and brain structures such as the hippocampus, prefrontal cortex and amygdala (1), making the brain more vulnerable to increased levels of GCs and metabolic changes and less resilient to damage from exposure to stressors (52,53); these alterations lead to formation of an abnormal aging brain with decreased function and neuronal death and stress serves a key role in brain aging.

GO analysis of upregulated DEGs in CUMS (6 weeks) rats showed functional alterations in ‘regulation of the apoptotic process’ and ‘negative regulation of cell proliferation’ and downregulated DEGs showed functional alterations in ‘receptor-mediated endocytosis’, ‘angiogenesis’, ‘synapse’, ‘dendrite’ and ‘axon’. This suggested that under chronic stress, neuronal apoptosis increased, multiple functions, such as cell proliferation and endocytosis, were inhibited and expression of genes associated with neuronal structures, such as synapses, axons and dendrites, decreased, resulting in damaging effects, such as impaired neuronal function and synaptic plasticity (46,49). Research has shown that under chronic stress, hippocampal dentate gyrus and CA3 area pyramidal cells atrophy and die (54), mossy fiber-CA3 synaptic transmission (55) dentate gyrus granule cell production decrease, dentate gyrus neurogenesis is inhibited (56) and CA3 area dendritic spine length and number are significantly decreased (57). Chronic stress causes functional and structural damage to hippocampal neurons, which exacerbates neuroendocrine abnormality and leads to mood changes and cognitive impairment (1). Parul et al (58) found that chronic unpredictable stress inhibits hippocampal neurogenesis in adult rats and increases glial cell and neuronal apoptosis in the cerebral cortex and hippocampus, which may be associated with reduced AKT and increased ERK signaling. Dagyte et al (59) concluded that strong activation of the HPA axis during acute stress is not sufficient to decrease hippocampal cell proliferation, but prolonged and repeated exposure to stressful stimuli inhibits hippocampal cell proliferation and leads to neurodevelopmental dysplasia. The present study supported the hypothesis that chronic stress may lead to cumulative changes in the brain that are not observed following acute stress. Such changes may indicate compromised brain plasticity and increased vulnerability to neuropathology (60).

In summary, genome-wide expression data and functional analysis reported here revealed dynamic functional changes during the different stages of CUMS. During the early and middle stages of chronic stress, the hippocampal ECM appears significantly altered and damage repair capacity is enhanced. In the late stage of CUMS, the balance of hippocampal apoptosis and proliferation is disrupted and chronic stress induces increased hippocampal apoptosis, aging, and the inhibition of multiple neuronal structures and functions. Further experiments are required to analyzed differential gene expression profiles and provide novel directions for gene therapy to treat chronic stress.

Supplementary Material

Supporting Data

Acknowledgements

Not applicable.

Funding

The present study was supported by the National Natural Science Foundation of China (grant nos. 31771290 and 81702454), National Natural Science Foundation of Beijing (grant no. 5222033) and Military Logistics Scientific Research Foundation of China (grant no. BWS17J027).

Availability of data and materials

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

Authors' contributions

LJQ, FX and FL contributed to the conception and design of the study. FL, FX, XW and YZ collected and interpreted the data. FL and FX confirm the authenticity of all the raw data. FL, YW and FX performed statistical analysis and wrote the manuscript. All authors contributed to manuscript revision. All authors have read and approved the manuscript.

Ethics approval and consent to participate

All experimental protocols were approved by the Institutional Animal Care and Use Committee of the Beijing Institute of Basic Medical Sciences (approval no. IACUC-DWZX-2020670) and were performed in accordance with the NIH Guide for the Care and Use of Laboratory Animals (NIH Publications no. 8023) and the ARRIVE guidelines from NC3Rs. During treatment, efforts were made to minimize the potential pain and distress of animals. Animals were monitored at least four times a week.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

McEwen BS, Bowles NP, Gray JD, Hill MN, Hunter RG, Karatsoreos IN and Nasca C: Mechanisms of stress in the brain. Nat Neurosci. 18:1353–1363. 2015. View Article : Google Scholar : PubMed/NCBI

2 

Weinberg A and Creed F: Stress and psychiatric disorder in healthcare professionals and hospital staff. Lancet. 355:533–537. 2000. View Article : Google Scholar : PubMed/NCBI

3 

Sinha R: Chronic stress, drug use, and vulnerability to addiction. Ann NY Acad Sci. 1141:105–130. 2008. View Article : Google Scholar : PubMed/NCBI

4 

Rozanski A, Blumenthal JA and Kaplan J: Impact of psychological factors on the pathogenesis of cardiovascular disease and implications for therapy. Circulation. 99:2192–2217. 1999. View Article : Google Scholar : PubMed/NCBI

5 

Bhatia V and Tandon RK: Stress and the gastrointestinal tract. J Gastroenterol Hepatol. 20:332–339. 2005. View Article : Google Scholar : PubMed/NCBI

6 

Reiche EM, Nunes SO and Morimoto HK: Stress, depression, the immune system, and cancer. Lancet Oncol. 5:617–625. 2004. View Article : Google Scholar : PubMed/NCBI

7 

Sapolsky RM: Stress and the brain: Individual variability and the inverted-U. Nat Neurosci. 18:1344–1346. 2015. View Article : Google Scholar : PubMed/NCBI

8 

Kim JJ and Diamond DM: The stressed hippocampus, synaptic plasticity and lost memories. Nat Rev Neurosci. 3:453–462. 2002. View Article : Google Scholar : PubMed/NCBI

9 

Kempermann G and Kronenberg G: Depressed new neurons-adult hippocampal neurogenesis and a cellular plasticity hypothesis of major depression. Biol Psychiatry. 54:499–503. 2003. View Article : Google Scholar : PubMed/NCBI

10 

Chao HM, Ma LY, McEwen BS and Sakai RR: Regulation of glucocorticoid receptor and mineralocorticoid receptor messenger ribonucleic acids by selective agonists in the rat hippocampus. Endocrinology. 139:1810–1814. 1998. View Article : Google Scholar : PubMed/NCBI

11 

Li XH, Chen JX, Yue GX, Liu YY, Zhao X, Guo XL, Liu Q, Jiang YM and Bai MH: Gene expression profile of the hippocampus of rats subjected to chronic immobilization stress. PLoS One. 8:e576212013. View Article : Google Scholar : PubMed/NCBI

12 

Ubaldi M, Ricciardelli E, Pasqualini L, Sannino G, Soverchia L, Ruggeri B, Falcinelli S, Renzi A, Ludka C, Ciccocioppo R and Hardiman G: Biomarkers of hippocampal gene expression in a mouse restraint chronic stress model. Pharmacogenomics. 16:471–482. 2015. View Article : Google Scholar : PubMed/NCBI

13 

Antoniuk S, Bijata M, Ponimaskin E and Wlodarczyk J: Chronic unpredictable mild stress for modeling depression in rodents: Meta-analysis of model reliability. Neurosci Biobehav Rev. 99:101–116. 2019. View Article : Google Scholar : PubMed/NCBI

14 

Xie F, Zhao Y, Ma J, Gong JB, Wang SD, Zhang L, Gao XJ and Qian LJ: The involvement of homocysteine in stress-induced Aβ precursor protein misprocessing and related cognitive decline in rats. Cell Stress Chaperones. 21:915–926. 2016. View Article : Google Scholar : PubMed/NCBI

15 

Bolger AM, Lohse M and Usadel B: Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics. 30:2114–2120. 2014. View Article : Google Scholar : PubMed/NCBI

16 

Kim D, Langmead B and Salzberg SL: HISAT: A fast spliced aligner with low memory requirements. Nat Methods. 12:357–360. 2015. View Article : Google Scholar : PubMed/NCBI

17 

Trapnell C, Williams BA, Pertea G, Mortazavi A, Kwan G, van Baren MJ, Salzberg SL, Wold BJ and Pachter L: Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat Biotechnol. 28:511–515. 2010. View Article : Google Scholar : PubMed/NCBI

18 

Anders S, Pyl PT and Huber W: HTSeq-a Python framework to work with high-throughput sequencing data. Bioinformatics. 31:166–169. 2015. View Article : Google Scholar : PubMed/NCBI

19 

Anders S and Huber W: Differential expression analysis for sequence count data. Genome Biol. 11:R1062010. View Article : Google Scholar : PubMed/NCBI

20 

Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, Davis AP, Dolinski K, Dwight SS, Eppig JT, et al: Gene ontology: Tool for the unification of biology. The gene ontology consortium. Nat Genet. 25:25–29. 2000. View Article : Google Scholar : PubMed/NCBI

21 

Pfaffl MW: A new mathematical model for relative quantification in real-time RT-PCR. Nucleic Acids Res. 29:e452001. View Article : Google Scholar : PubMed/NCBI

22 

Du J, Wang Y, Hunter R, Wei Y, Blumenthal R, Falke C, Khairova R, Zhou R, Yuan P, Machado-Vieira R, et al: Dynamic regulation of mitochondrial function by glucocorticoids. Proc Natl Acad Sci USA. 106:3543–3548. 2009. View Article : Google Scholar : PubMed/NCBI

23 

Daskalakis NP, Cohen H, Cai G, Buxbaum JD and Yehuda R: Expression profiling associates blood and brain glucocorticoid receptor signaling with trauma-related individual differences in both sexes. Proc Natl Acad Sci USA. 111:13529–13534. 2014. View Article : Google Scholar : PubMed/NCBI

24 

Govindarajan A, Rao BS, Nair D, Trinh M, Mawjee N, Tonegawa S and Chattarji S: Transgenic brain-derived neurotrophic factor expression causes both anxiogenic and antidepressant effects. Proc Natl Acad Sci USA. 103:13208–13213. 2006. View Article : Google Scholar : PubMed/NCBI

25 

Lakshminarasimhan H and Chattarji S: Stress leads to contrasting effects on the levels of brain derived neurotrophic factor in the hippocampus and amygdala. PLoS One. 7:e304812012. View Article : Google Scholar : PubMed/NCBI

26 

Joëls M and Baram TZ: The neuro-symphony of stress. Nat Rev Neurosci. 10:459–466. 2009. View Article : Google Scholar : PubMed/NCBI

27 

Pu Z, Krugers HJ and Joëls M: Corticosterone time-dependently modulates beta-adrenergic effects on long-term potentiation in the hippocampal dentate gyrus. Learn Mem. 14:359–367. 2007. View Article : Google Scholar : PubMed/NCBI

28 

Clark WC, Yang JC and Janal MN: Altered pain and visual sensitivity in humans: The effects of acute and chronic stress. Ann NY Acad Sci. 467:116–129. 1986. View Article : Google Scholar : PubMed/NCBI

29 

Tijerina L, Garrott WR, Stoltzfus D and Parmer E: Eye glance behavior of van and passenger car drivers during lane change decision phase. Transp Res Rec. 1937:37–43. 2005. View Article : Google Scholar

30 

Paul M, Lech RK, Scheil J, Dierolf AM, Suchan B and Wolf OT: Acute stress influences the discrimination of complex scenes and complex faces in young healthy men. Psychoneuroendocrinology. 66:125–129. 2016. View Article : Google Scholar : PubMed/NCBI

31 

Smith KE, Leitzke BT and Pollak SD: Youths' processing of emotion information: Responses to chronic and video-based laboratory stress. Psychoneuroendocrinology. 122:1048732020. View Article : Google Scholar : PubMed/NCBI

32 

Norton DJ, McBain RK, Pizzagalli DA, Cronin-Golomb A and Chen Y: Dysregulation of visual motion inhibition in major depression. Psychiatry Res. 240:214–221. 2016. View Article : Google Scholar : PubMed/NCBI

33 

Golomb JD, McDavitt JR, Ruf BM, Chen JI, Saricicek A, Maloney KH, Hu J, Chun MM and Bhagwagar Z: Enhanced visual motion perception in major depressive disorder. J Neurosci. 29:9072–9077. 2009. View Article : Google Scholar : PubMed/NCBI

34 

Friberg TR and Borrero G: Diminished perception of ambient light: A symptom of clinical depression? J Affect Disord. 61:113–118. 2000. View Article : Google Scholar : PubMed/NCBI

35 

Anwar MM, Özkan E and Gürsoy-Özdemir Y: The role of extracellular matrix alterations in mediating astrocyte damage and pericyte dysfunction in Alzheimer's disease: A comprehensive review. Eur J Neurosci. Jun 28–2021.(Epub ahead of print). View Article : Google Scholar : PubMed/NCBI

36 

Rosenberg GA: Extracellular matrix inflammation in vascular cognitive impairment and dementia. Clin Sci (Lond). 131:425–437. 2017. View Article : Google Scholar : PubMed/NCBI

37 

Hornberger LK, Singhroy S, Cavalle-Garrido T, Tsang W, Keeley F and Rabinovitch M: Synthesis of extracellular matrix and adhesion through beta(1) integrins are critical for fetal ventricular myocyte proliferation. Circ Res. 87:508–515. 2000. View Article : Google Scholar : PubMed/NCBI

38 

Henriet P, Zhong ZD, Brooks PC, Weinberg KI and DeClerck YA: Contact with fibrillar collagen inhibits melanoma cell proliferation by up-regulating p27KIP1. Proc Natl Acad Sci USA. 97:10026–10031. 2000. View Article : Google Scholar : PubMed/NCBI

39 

Kelly KK, MacPherson AM, Grewal H, Strnad F, Jones JW, Yu J, Pierzchalski K, Kane MA, Herson PS and Siegenthaler JA: Col1a1+ perivascular cells in the brain are a source of retinoic acid following stroke. BMC Neurosci. 17:492016. View Article : Google Scholar : PubMed/NCBI

40 

Falconer MA, Serafetinides EA and Corsellis JA: Etiology and pathogenesis of temporal lobe epilepsy. Arch Neurol. 10:233–248. 1964. View Article : Google Scholar : PubMed/NCBI

41 

Thom M: Review: Hippocampal sclerosis in epilepsy: A neuropathology review. Neuropathol Appl Neurobiol. 40:520–543. 2014. View Article : Google Scholar : PubMed/NCBI

42 

Rosensteel SM, Wilson RP, White SL and Ehrlich HP: COL1A1 oligodeoxynucleotides decoy: Biochemical and morphologic effects in an acute wound repair model. Exp Mol Pathol. 89:307–313. 2010. View Article : Google Scholar : PubMed/NCBI

43 

Shi-Wen X, Leask A and Abraham D: Regulation and function of connective tissue growth factor/CCN2 in tissue repair, scarring and fibrosis. Cytokine Growth Factor Rev. 19:133–144. 2008. View Article : Google Scholar : PubMed/NCBI

44 

Krishnaswamy VR, Benbenishty A, Blinder P and Sagi I: Demystifying the extracellular matrix and its proteolytic remodeling in the brain: Structural and functional insights. Cell Mol Life Sci. 76:3229–3248. 2019. View Article : Google Scholar : PubMed/NCBI

45 

Moreno-Fernandez RD, Tabbai S, Castilla-Ortega E, Perez-Martin M, Estivill-Torrus G, Rodriguez de Fonseca F, Santin LJ and Pedraza C: Stress, depression, resilience and ageing: A role for the LPA-LPA1 pathway. Curr Neuropharmacol. 16:271–283. 2018. View Article : Google Scholar : PubMed/NCBI

46 

de Magalhães JP and Passos JF: Stress, cell senescence and organismal ageing. Mech Ageing Dev. 170:2–9. 2018. View Article : Google Scholar : PubMed/NCBI

47 

Segar TM, Kasckow JW, Welge JA and Herman JP: Heterogeneity of neuroendocrine stress responses in aging rat strains. Physiol Behav. 96:6–11. 2009. View Article : Google Scholar : PubMed/NCBI

48 

Sapolsky RM, Krey LC and McEwen BS: The neuroendocrinology of stress and aging: The glucocorticoid cascade hypothesis. Endocr Rev. 7:284–301. 1986. View Article : Google Scholar : PubMed/NCBI

49 

Raber J: Detrimental effects of chronic hypothalamic-pituitary-adrenal axis activation. From obesity to memory deficits. Mol Neurobiol. 18:1–22. 1998. View Article : Google Scholar : PubMed/NCBI

50 

Elgh E, Lindqvist Astot A, Fagerlund M, Eriksson S, Olsson T and Näsman B: Cognitive dysfunction, hippocampal atrophy and glucocorticoid feedback in Alzheimer's disease. Biol Psychiatry. 59:155–161. 2006. View Article : Google Scholar : PubMed/NCBI

51 

Buwembo A, Long H and Walker CD: Participation of endocannabinoids in rapid suppression of stress responses by glucocorticoids in neonates. Neuroscience. 249:154–161. 2013. View Article : Google Scholar : PubMed/NCBI

52 

Conrad CD: Chronic stress-induced hippocampal vulnerability: The glucocorticoid vulnerability hypothesis. Rev Neurosci. 19:395–411. 2008. View Article : Google Scholar : PubMed/NCBI

53 

Bloss EB, Janssen WG, McEwen BS and Morrison JH: Interactive effects of stress and aging on structural plasticity in the prefrontal cortex. J Neurosci. 30:6726–6731. 2010. View Article : Google Scholar : PubMed/NCBI

54 

Magariños AM, McEwen BS, Flügge G and Fuchs E: Chronic psychosocial stress causes apical dendritic atrophy of hippocampal CA3 pyramidal neurons in subordinate tree shrews. J Neurosci. 16:3534–3540. 1996. View Article : Google Scholar : PubMed/NCBI

55 

Magariños AM, Verdugo JM and McEwen BS: Chronic stress alters synaptic terminal structure in hippocampus. Proc Natl Acad Sci USA. 94:14002–14008. 1997. View Article : Google Scholar : PubMed/NCBI

56 

Gould E and Tanapat P: Stress and hippocampal neurogenesis. Biol Psychiatry. 46:1472–1479. 1999. View Article : Google Scholar : PubMed/NCBI

57 

McEwen BS: Stress and hippocampal plasticity. Annu Rev Neurosci. 22:105–122. 1999. View Article : Google Scholar : PubMed/NCBI

58 

Parul Mishra A, Singh S, Singh S, Tiwari V, Chaturvedi S, Wahajuddin M, Palit G and Shukla S: Chronic unpredictable stress negatively regulates hippocampal neurogenesis and promote anxious depression-like behavior via upregulating apoptosis and inflammatory signals in adult rats. Brain Res Bull. 172:164–179. 2021. View Article : Google Scholar : PubMed/NCBI

59 

Dagyte G, Van der Zee EA, Postema F, Luiten PG, Den Boer JA, Trentani A and Meerlo P: Chronic but not acute foot-shock stress leads to temporary suppression of cell proliferation in rat hippocampus. Neuroscience. 162:904–913. 2009. View Article : Google Scholar : PubMed/NCBI

60 

McEwen BS and Morrison JH: The brain on stress: Vulnerability and plasticity of the prefrontal cortex over the life course. Neuron. 79:16–29. 2013. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

April-2022
Volume 25 Issue 4

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 F, Wang Y, Wang X, Zhao Y, Xie F and Qian L: Dynamic effects of chronic unpredictable mild stress on the hippocampal transcriptome in rats. Mol Med Rep 25: 110, 2022
APA
Li, F., Wang, Y., Wang, X., Zhao, Y., Xie, F., & Qian, L. (2022). Dynamic effects of chronic unpredictable mild stress on the hippocampal transcriptome in rats. Molecular Medicine Reports, 25, 110. https://doi.org/10.3892/mmr.2022.12626
MLA
Li, F., Wang, Y., Wang, X., Zhao, Y., Xie, F., Qian, L."Dynamic effects of chronic unpredictable mild stress on the hippocampal transcriptome in rats". Molecular Medicine Reports 25.4 (2022): 110.
Chicago
Li, F., Wang, Y., Wang, X., Zhao, Y., Xie, F., Qian, L."Dynamic effects of chronic unpredictable mild stress on the hippocampal transcriptome in rats". Molecular Medicine Reports 25, no. 4 (2022): 110. https://doi.org/10.3892/mmr.2022.12626