• Open Access
    Review

    Cellular and molecular mechanisms of stress-induced memory impairment

    Ameneh Rezayof 1*
    Maryam Sardari 1
    Shiva Hashemizadeh 2

    Explor Neurosci. 2022;1:100–119 DOI: https://doi.org/10.37349/en.2022.00008

    Received: June 29, 2022 Accepted: September 28, 2022 Published: December 30, 2022

    Academic Editor: Ertugrul Kilic, Istanbul Medipol University, Turkey

    This article belongs to the special issue Neuroinflammation in the Ageing and the Injured Brain

    Abstract

    Exposure to stressful conditions plays a critical role in brain processes, including neural plasticity, synaptic transmission, and cognitive functions. Since memory-related brain regions, the hippocampus (Hip), the amygdala, and the prefrontal cortex, express high glucocorticoid receptors (GRs), these areas are the potential targets of stress hormones. Stress affects memory encoding, consolidation, and retrieval, which may depend on many factors such as the type, duration, the intensity of the stressor or the brain region. Here, this review mainly focused on the mechanisms involved in stress-induced memory impairment. Acute/chronic stress induces structural and functional changes in neurons and glial cells. Dendritic arborization, reduction of dendritic spine density, and alteration in glutamatergic-mediated synaptic transmission via N-methyl-D-aspartate (NMDA) and α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA) receptors are mechanisms that stress affect long-term memory formation. Exposure to acute or chronic stress could interplay with multiple neurotransmitter signaling, modulating the neuronal circuits involved in memory impairment or state-dependent learning. Stress hormones also modulate the expression of microRNAs in the specific brain regions responsible for stress-induced behaviors. Because of expressing GRs in astrocytes and microglial cells, stress could affect the morphology, structure, and functions of these glial cells in memory-related brain regions. Astrocytes play a crucial role in stress-induced aversive or fear memory formation. Over-activation of the microglial cells enhances the release of inflammatory cytokines, which results in neuronal injury. Stress has a prominent role in cognitive decline to induces memory problems, particularly in older adults. Due to the issue’s importance, here the provided overview attempted to address the question of how stress alters neuronal epigenetic regulators, synaptic transmissions, and glial activity in the brain.

    Keywords

    Stress, memory impairment, neurotransmitters, astrocytes, microglia, microRNAs

    Introduction

    Stress is an adaptive neurobiological response that enables the body to adjust to internal and external environmental challenges [1]. Stress-responsive system is a complex adaptive system that seems to be highly conserved across species [2]. It is well established that two different systems but interconnected, including the hypothalamic-pituitary-adrenocortical (HPA) and the sympathetic-adrenomedullary (SAM), have significant roles in the coordination of both physical and psychological responses to stress situations [3]. The SAM system is a component of the sympathetic division of the autonomic nervous system. It is responsible for the first phase of stress response which leads to a short-term reaction to provide appropriate responses [4]. Following the activation of the SAM system, the adrenal gland center releases epinephrine (adrenaline) into the blood circulation. Then, the increased circulating epinephrine facilitates a rapid mobilization of metabolic resources to induce fight/flight fast responses [5, 6]. Exposure to a stressor increases the secretion of corticotropin-releasing hormone (CRH) and arginine vasopressin (AVP) from the paraventricular nucleus (PVN) of the hypothalamus to stimulate the anterior pituitary gland to secrete adrenocorticotropic hormone (ACTH) into the blood circulation [7, 8]. The ACTH, in turn, induces glucocorticoid synthesis and release from the cortex of adrenal glands, which are cortisol in humans and corticosterone in rodents [9]. Evidence suggests that endogenous synthesis of corticosteroids is not limited to the adrenal cortex. Using messenger RNA (mRNA) and immunogold electron microscopic analysis, it has been shown that the hippocampal neurons express the specific enzymes, including P450 (c21), P450 (2D4), P450 (11β1), and 3β-hydroxysteroid dehydrogenase to produce a low nanomolar level of corticosterone in rats [10].

    The glucocorticoids are responsible for long-term responses and help the body adapt to stress. Circulating glucocorticoids seems necessary for brain development, neuronal survival and neurogenesis, psychophysiological adaptation to stress conditions, and adaptive immune responses [11, 12]. The secretion of glucocorticoids from the HPA axis is tightly regulated by endocrine and neuronal systems [13]. The rate of glucocorticoid release and the level of glucocorticoid release could determine ‘fast’ and ‘delayed’ negative feedback regulation, respectively, to protect the body from prolonged activities [14]. Glucocorticoid negative feedback may be mediated at multiple levels via glucocorticoid receptors (GRs) located in the hypothalamus and the anterior pituitary gland. The HPA axis is also modulated through a glucocorticoid-independent neuronal mechanism [13]. Both direct and indirect pathways have a central role in controlling the end product of the HPA axis (Figure 1) [15].

    Stress affects the hippocampus (Hip), the amygdala (AMY), the prefrontal cortex (PFC), and the HPA axis. Following stress exposure, the hypothalamus is activated to release CRH into the bloodstream to stimulate the pituitary gland for producing ACTH. The bloodstream delivers ACTH to the adrenal glands to release cortisol in humans and corticosterone in rodents. Stress hormones via GRs affect memory-related brain regions. +: stimulation; –: inhibition

    The limbic brain structures, including the Hip, the AMY, and the PFC, control the PVN activity via the glutamatergic and gamma aminobutyric acidergic (GABAergic) innervations [16]. Glutamatergic projections of the Hip and the PFC indirectly inhibit the PVN neurons in the hypothalamus. Thus, they play an essential role in terminating the HPA axis-induced stress response and have a negative feedback modulation on this axis [17, 18]. Neuroimaging studies have shown that the decrease in the anterior cingulate cortex volume could be associated with the dysregulation of HPA activity [19, 20]. Electrical stimulation of the Hip decreased glucocorticoid secretion [21], whereas the lesion of this brain region caused the stress response [22]. In contrast, the AMY GABAergic projection disinhibits the hypothalamic PVN neurons resulting in positive regulation of the HPA activity [23]. AMY activation is associated with stress-related behaviors in an emotional state, such as fear [24].

    Corticosteroid hormones affect neurons and glia through two intracellular receptors: type 1 mineralocorticoid (MRs) and type 2 GRs. They are members of the nuclear receptor superfamily of ligand-dependent transcription factors [25]. Since MRs have a higher affinity to glucocorticoids than GRs, it is likely that type 1 receptors are primarily occupied in basal/non-stressful conditions to maintain stress responses. On the other hand, the activation of GRs through a high level of glucocorticoids following stress exposure is associated with the neuroendocrine stress response to restore homeostasis by a negative feedback loop [26]. The different distribution of GRs and MRs in the brain regions, including the hypothalamus [27], the PFC [28], the Hip [11], and the AMY [29] have directly been targeted by psychogenic and physical stressors. Interestingly, stress exposure alters mRNA expressions of both MRs and GRs in the brain regions, including the Hip and the AMY. The exposure to single prolonged stress (SPS) caused an early decrease in GR and MR mRNAs and protein expressions [29]. Evidence suggests that both acute and chronic stress could alter the expression levels of GR in the PFC, the Hip, and the hypothalamus to modulate GR phosphorylation [30]. Hence, these brain regions may be targeted directly by psychogenic and physical stressors.

    Exposure to stressful conditions has a significant influence on learning and memory processes. Memory formation is associated with neuronal synaptic changes [31], significantly strengthening existing synapses by increasing the number and size of the dendritic spine with concomitant changes in synaptic markers. Although the general outcome of stress is memory impairment [32], some studies have reported the neutral [33] or even the facilitative role of stress on memory formation [34]. It seems that depending on the type, intensity, and duration of stressors, the stress has a variety of effects on memory formation. In this review, the main focus was on the mechanisms involved in stress-induced memory impairment.

    Effects of stress on memory

    Stress impairs memory formation

    The end product of neuroendocrine stress cascades, glucocorticoids, plays a critical role in brain processes, including neural plasticity, synaptic transmission, and cognitive functions such as learning and memory. Glucocorticoids exert their cellular effects by acting on MRs, and GRs in the brain regions, which are essential for memory performance [35, 36]. The elevation of cortisol under a stressful situation or exogenous glucocorticoid administration affects memory consolidation. The exposure to the acute stressful stimulus affected the consolidation of object recognition task memory in mice and resulted in long-term memory impairment without any adverse effect on short-term memory. Blocking cannabinoid receptors (CBs) of adrenergic neurons improved stress-induced memory loss [37], indicating the role of the cannabinoid system in the modulation of cognitive function. In line with these findings, human research has shown that acute stress in pre-encoding information enhances the recall of emotional memories [38]. Moreover, an investigation has revealed an inverted U-shape curve relation between stress intensity and memory acquisition and reconsolidation [39]. On the other hand, memory impairment is dependent on cortisol level, for example, exposure to acute stress before long-term retrieval leads to loss of memory recall at the lowest cortisol level. In contrast, a concurrent rise in cortisol levels enhanced long-term memory [40]. Long-term memory was negatively affected by acute stress during memory consolidation and before memory reactivation. In a study conducted by Sardari et al. [32], it has been shown that exposure to inescapable stress before the retention trial of a passive avoidance task induces memory retrieval impairment. Furthermore, post-training exposure to stress also decreased memory retrieval on the test day [34, 41]. Interestingly, before being exposed to the retention stage, the same stressful situation completely reversed stress-induced memory impairment and showed a typical memory performance [34]. This result indicated that recalling a memory under stress is state-dependent, and memory enhancement occurs in a similar stressful state that happens during the consolidation of information.

    Stress-induced robust changes in structural and functional neuronal plasticity [42] eventually altered glutamatergic-mediated synaptic transmission via N-methyl-D-aspartate (NMDA) and α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA) receptors, which are essential components in synaptic plasticity, therefore, they could affect memory storage [43]. Experiments prove that stress changes dendritic arborization and alters dendritic spine density in different brain regions. Chronic stress results in spine density loss and retraction of apical dendritic branches in the hippocampal CA1 and CA3 regions [44]. This effect was reversed by the CRH receptor type 1 (CRHR1) antagonist [44, 45]. The activation of NMDA receptor-mediated CRH-induced dendritic spine loss through calpain activation was shown previously [46]. Furthermore, the PFC, a cortical area important for working memory, executive functioning, and goal-directed behaviors, has a most vulnerable dendritic domain when interacting with stressful events. Moreover, acute stress potentiates glutamatergic transmission in the PFC [47], facilitating working memory through glucocorticoid-inducible kinase signaling [48]. In contrast, chronic stress causes the impairment of prefrontal-dependent memory formation [49] by acting on suppressing glutamatergic transmission. The effects of stress on dendritic remodeling are region-specific. Inactivation of the basolateral AMY (BLA) before chronic stressors prevents volume reduction of the PFC [50].

    Stress produces state-dependent learning

    State-dependent learning refers to a special kind of learning in which the retrieval of an event requires the animal to be in a similar state of acquired information [51]. Our previous study has indicated that exposure to acute elevated platform stress has an amnesic effect on memory consolidation and retrieval in the passive avoidance task. Our results also indicated that pre-test exposure to stress reversed the post-training stress-induced memory impairment. Since the information retrieval was performed if the animal was in the same stress state during the encoding phase, our data indicated that acute stress produces state-dependent memory retrieval (Figure 2) [34]. An autoradiographic study also showed that re-exposure to acute restraint stress enhanced the AMY NMDA receptor activity [52]. Intra-BLA microinjection of a selective antagonist of 5-hydroxytryptamine 1A receptor (5-HT1A) receptors inhibited state-dependent learning under stress in rats. Several mechanisms seem involved in stress-induced state-dependent learning, and further studies should figure out the mechanisms.

    Effect of stress on passive avoidance memory. A: Post-training exposure to acute elevated platform stress (30 min) impaired memory consolidation in rats. B: Pre-test exposure to the same acute stress-induced memory retrieval impairment. A/C: The memory impairment induced by post-training exposure to stress was restored in the animals that received same pre-test acute stress exposure named stress-induced state-dependent memory retrieval

    Stress affects long-term potentiation and long-term depression

    The activation of the HPA axis and SAM system following exposure to a stressful event could modulate cognitive and emotional memory through changes in synaptic strength [53]. Some studies reported that stress induces alterations in synaptic plasticity through cellular mechanisms, including long-term potentiation (LTP) and long-term depression (LTD). Electrophysiological recordings were implicit that stress has a distinct effect on the efficacy of glutamatergic synaptic transmission in the Hip, the PFC, and the AMY [54]. For example, high glucocorticoid levels induced NMDA-dependent LTP impairment in the CA1 region and did not affect the dentate gyrus (DG). Acute stress attenuated LTP formation in the dorsal Hip, while it enhanced LTP in the ventral Hip [5557]. Acute stress impaired mGlu3-LTD in the specific synapse between the BLA and the PFC, which may be correlated with the cognitive dysfunction of the PFC during stress-related disorders [58]. Unpredictable shock stress has been shown to have no significant effect on spine morphology but could enhance synaptic plasticity in principal the BLA neurons [59]. A CB1 receptor antagonist could recover chronic stress-induced LTP impairment, furthermore, repeated stress could alter muscarinic LTP in the hippocampal slices [60]. Moreover, chronic stress concurrent with exercise could improve the dorsal hippocampal LTP associated with an increased expression of BDNF to enhance memory [61].

    The neuro-cognitive processes contain multicomponent stages, including encoding, consolidation, and retrieval. The most intensive studies showed that stress might affect all three specific phases involved in memory formation and directly is dependent on many factors such as the stressor’s timing, which in turn, cause improvement or impairment of memory performance [35]. The animal examination has indicated that elevation of cortisol under stressful situations or exogenous glucocorticoid administration affects memory consolidation. The exposure to the acute stressful stimulus affected the consolidation of object recognition memory in mice and resulted in long-term memory impairment without any adverse effect on short-term memory. Interestingly stress-induced memory loss is reversed by blocking CBs of adrenergic neurons [37], indicating the role of the cannabinoid system in the modulation of cognitive function. In line with these findings, human research has shown that acute stress in pre-encoding information enhances the recall of emotional memories [38]. An inverted U-shape curve relation exists between stress intensity and memory acquisition and reconsolidation [39]. Memory impairment may depend on cortisol level. For example, the exposure to acute stress before the long-term retrieval led to the loss of memory recall in the lowest cortisol level, while long-term memory was enhanced with a concurrent rise in cortisol level [40].

    Long-term memory was negatively affected by acute stress during memory consolidation and before memory reactivation. Sardari et al. [32] showed that exposure to inescapable stress before the retention trial of a passive avoidance task induces memory retrieval impairment. Furthermore, post-training exposure to stress decreased memory retrieval on the test day [34, 41]. The same stress situation just before exposure to the retention stage completely reversed stress-induced memory impairment [34]. These findings indicated that memory recall is state-dependent, and memory enhancement occurs in a similar stressful state during the consolidation of information.

    Neurotransmitters are involved in the effects of stress on memory

    Different stressors affect various neurotransmitter systems’ molecular and cellular functions, including GABAergic, cholinergic, glutamatergic, serotonergic, dopaminergic, and endocannabinoid system (ECS). The changes in neurotransmitters’ release and synaptic concentrations are associated with the type of stress and brain regions [62]. A large body of evidence considers the effects of stress on neurotransmitter systems in learning and memory processes. Here, we provide an overview of how stress-induced neurotransmission changes may have happened during the encoding and retaining of new information in memory-related brain regions.

    GABAergic system

    GABAergic dysfunction leads to neurological and mental diseases [63, 64]. Gamma-aminobutyric acid (GABA) receptors, including GABA subtype A (GABAA), GABA subtype B (GABAB), and GABA subtype C (GABAC), are ionotropic receptors that control and manage cognitive functions broadly expressed in the central nervous system (CNS). Metabotropic GABAB receptors bind to G protein inhibitors (Gi/o) to mediate prolonged slow inhibitory action. Both pre- and post-synaptic GABAA and GABAB receptors inhibit the excitatory and inhibitory neuronal functions. Hence, activation of these receptors modulates LTP in the different brain regions [65, 66]. Stressful conditions alter the GABAergic transmission to change and adapt emotional and behavioral responses. For example, exposure to new environmental stressors or swimming stress increased GABA release in the Hip of mice [67]. Acute restraint stress increased GABA efflux in the BLA. In contrast, the same type of stress did not affect the central AMY efflux using an in vivo microdialysis technique [68].

    The GABAergic system modulates the HPA axis mainly through GABAA receptors. In normal conditions, CRH neurons of the PVN are under tonic inhibition of GABA transmission. Stressful situations enhance the release of GABA and activation of AMY GABAA receptors, eliminating the tonic inhibition of GABA to increase the release rate of stress hormones [69]. Our previous results showed that inhibiting GABAA receptors in the BLA via microinjection of muscimol, a GABAA receptor agonist, increased the response of ineffective acute stress to impair memory retrieval in a passive avoidance paradigm [32]. Interestingly, prenatal stress decreased the hippocampal spine densities and impaired spatial memory formation in adult offspring by enhancing the GABA transmission in the brain’s developmental stage [70]. Therefore, GABA receptors may play a critical role in regulating the activity of the HPA axis to mediate memory formation under stress.

    Dopaminergic system

    Dopamine-induced effects are mediated via G-protein coupled receptors classified into two subclasses: D1- and D2-like receptors. Stimulation of dopaminergic D1 or D2 receptors enhances or inhibits the adenylyl cyclase activity to change cyclic adenosine monophosphate (cAMP) concentration, respectively [71]. Dopamine receptors are highly expressed in different brain areas, including the ventral tegmental area (VTA), the nucleus accumbens (NAc), the substantia nigra, the Hip, and the AMY [72]. Dopamine receptors are involved in the formation of different types of learning and memory, such as passive avoidance learning [73], spatial memory [74], and reward-related learning and memory [75].

    Exposure to acute or chronic stress increases dopamine release in the NAc, the VTA, and the PFC [76]. In general, acute stress increases the amount of dopamine in the brain to increase the risk of an adult’s desire to use addictive substances [77]. Blocking the PFC D1 receptors attenuated the conditioned place preference (CPP) induced by immobility stress in rats [78]. Additionally, inhibition of the PFC D1 receptors reversed the stress-induced working memory impairment [79]. There is increasing concern that the activation of these receptors may reduce the glutamate release and thereby induces the reduction of neuronal firing rate in the PFC [80]. Systemic administration of sulpiride, a selective D2 receptor antagonist, attenuated glucocorticoid-induced memory retrieval impairment [81]. In contrast, our previous study also showed that inactivation of the BLA D1 or D2 receptors inhibited the improving effects of nicotine on stress-induced memory retention impairment (Figure 3) [82]. These findings suggest that the dopamine receptors alone or in cooperation with other neurotransmitters modulate the impressive effects of stress hormones on memory formation.

    Effects of pre-test intra-BLA microinjections of D1 or D2 receptors antagonists on nicotine-induced improvement of memory retrieval impairment under stress. Pre-test exposure to acute elevated platform stress (30 min) impaired memory retrieval in rats. Nicotine administration improved stress-induced memory retrieval impairment. Pre-test intra-BLA microinjection of SCH23390 (a selective dopamine D1 receptor antagonist) or sulpiride (a selective dopamine D2 receptor antagonist) inhibited nicotine-induced improvement of the stress amnesic effect. s.c.: subcutaneous

    Serotonergic system

    Serotonin is mainly produced in the raphe nuclei of the brain stem. There are seven types of serotonergic receptors. Based on the chemical structure of serotonin, these receptors are called 5-hydroxytryptamine (5-HT). The serotonergic receptors are a combination of metabotropic and ionotropic receptors that have excitatory and inhibitory roles in the CNS [83]. The 5-HT receptors regulate the release of different neurotransmitters, including glutamate, GABA, and acetylcholine. The serotonergic heteroreceptors are therapeutic targets for treating depression, Alzheimer’s diseases (AD), and Parkinson’s diseases [84]. DNA microarray analysis was used to show that memory formation increases the differential gene expression of 5-HT receptors. For example, there is an enhancement of gene expression of 5-HT1A-1F, 5-HT2A, and 5-HT5A receptors during the passive avoidance memory formation. By contrast, spatial memory requires the gene expression of 5-HT2C, 5-HT3A, and 5-HT6 receptors in the Morris water maze [85].

    Evidence suggests that exposure to chronic stress leads to depression and anxiety disorders [86, 87]. Chronic foot shock stress caused depression, which was reversed by injecting antidepressants and increasing serotonin levels in the mice’s brains [88]. Antidepressants attenuate stress-related mood disorders and efficiently improve stress-induced amnesia. Recently, Léa Blondelle et al. [89] showed that exposure to mild unpredictable stress impaired spatial memory formation while using Bombax costatum as an antidepressant attenuated the amnesic effect of stress. Since serotonin plays a role in chronic stress-induced cognitive dysfunction, Natarajan and coworkers [90] found that exposure to chronic stress caused cell death in the interfascicular nucleus of the dorsal raphe, which results in decreased serotonergic innervation of the medial PFC (mPFC). However, they showed that the treatment with MK801, a competitive NMDA receptor antagonist, blocked stress-induced deficits in memory recall. Thus, one may suggest that chronic stress may increase glutamate release, which results in serotonergic neuronal injury in raphe nuclei [90]. Exposure to chronic stress enhanced the AMY sensitivity to serotonin, and the blockade of 5-HT2C receptors attenuated the stress-related fear memory in mice [91]. Inactivation of the ventral hippocampal 5-HT7 receptors also reversed the stress-induced freezing behavior and fear memory formation [92]. Our previous study showed that intra-BLA microinjection of (S)-WAY-100135, a selective antagonist of 5-HT1A receptors, prevented the impairing effect of stress on memory consolidation and retrieval in the passive avoidance learning task [34].

    ECS

    The ECS comprises a neuromodulatory system to mediate synaptic plasticity and neurogenesis, which are essential for memory formation. Anandamide and 2-arachidonoylglycerol (2-AG) as arachidonic acid (AA) derivatives bind to the CBs, which are densely expressed throughout our brains and bodies. CBs are metabotropic receptors that bind to inhibitory G-protein, and activating these receptors inhibits neuronal adenyl cyclase activity. Three types of CBs have been identified in the brain known as CB1, CB2, and G protein-coupled receptor 55 (GRP55). CB1 receptors are abundantly expressed in the CNS, including the AMY and the Hip. They are mainly responsible for mediating the effects of endocannabinoids in the brain [93]. Due to the high expression of CB1 receptors on presynaptic terminals, these receptors act as neuromodulators to inhibit neurotransmitter release [94]. Generally, activation of CB1 receptors leads to the impairment of hippocampal-related memory [95, 96]. Consumption of marijuana impairs the acquisition, consolidation, and retrieval of memory formation in humans [97]. On the other hand, activating the BLA CB1 receptors facilitates fear memory formation [98]. Cannabinoids have facilitating or inhibitory effects on memory formation based on the type of memory, brain regions, and memory stages [99, 100]. Notably, under pathological conditions, activation of the CBs has neuroprotective effects [101]. For instants, the activation of the CB1 receptor improved memory formation in AD animal models with memory deficits and cognitive disorders [102, 103]. Regarding stress-related disorders, it seems that cannabinoids improve stress-induced anxiety [99, 104]. Since chronic stress impairs cognitive function, the ECS may have a modulatory role in glucocorticoid-mediated outcomes [105]. Exposure to acute stress after memory consolidation leads to impairment of long-term object recognition. Intra-hippocampal administration of a CB antagonist prevented memory deficit [37]. Moreover, following footshock stress, activation of CBs abolished memory loss and facilitated LTP in the Hip [106]. Enhancement of cannabinoid signaling through local administration of a CB1/CB2 receptors agonist, WIN 55,212-2, into the dorsal striatum increased memory consolidation in a passive avoidance task. Moreover, central or peripheral blockade of CB1 receptor signaling eliminated the effect of acute stress-induced memory enhancement [107]. Collectively, these findings demonstrated that cannabinoid signaling pathways might serve as a potential therapeutic target to regulate glucocorticoid-mediated stress memory performance.

    microRNAs mediate memory formation under stress

    In 2001, the presence of microRNAs (miRNAs, small endogenous RNAs) in the mammalian system was reported for the first time. Currently, it is well established that these non-coding RNAs are involved in the various fundamental biological processes, including development concerning their role in fine-tuning gene expression based on post-transcriptional and translational regulation [108, 109]. Therefore, the function of miRNAs might correspond to the onset and pathophysiology of various neurobiological diseases [110].

    More recently, the alternative expression levels of miRNAs in the brain regions were suggested to be associated with the behavioral stress response. In a study conducted by Volk et al. [111], it has been shown that chronic stress increases the level of miR-15a in the AMY, which in turn decreases FK506-binding protein 51 (FKBP51) as a specific transcription factor for GR activity. Therefore, one may suggest that miR-15a is essential for stress adaptation in chronic stress. Another experiment implicated that foot shock stress could upregulate miR-34 in the adolescent rat hypothalamus and correlated with the decrease of CRF receptor type 1 (CRFR1) expression as a validated target for miR-34 [112, 113]. Moreover, the upregulation of miR-34 inhibits dendritic spine growth, which is associated with memory loss [114]. Furthermore, anxiety-like behavior in response to acute stress was reduced in miR-34 knockout mice. It seems that a lack of miR-34, a critical modulator of the stress response, could protect from the aversive stress effect [115].

    Other evidence supports that exposure to early life stress induces a change in the expression of clusters of miRNAs which contribute to memory impairment. Liu et al. [116] determined that spatial memory was impaired in response to early stress, and the expression patterns of miR-135a and miR-16 were changed in rats. They reported that the miR-135a level was reduced in the PFC, while miR-16 was upregulated in the rat Hip [116]. Both miRNAs have been associated with the regulation of serotonin levels in the synaptic cleft. For instance, the downregulation of miR-135 leads to declining serotonin levels and corresponds to stress-induced behavior [117]. Moreover, acute stress downregulates brain-specific miR-135a and miR-124 in the mice AMY and positively increases the expression of the protein level of the GR (MR), therefore acts as a mediator of stress response in the AMY [118].

    Recent studies revealed that early transient changes of the miRNA hippocampal expressions without concomitant alteration on the AMY miRNAs are necessary for Hip-dependent fear memory. Fear memory is closely related to stress and affects the brain structure related to learning and memory processes. Moreover, early up-regulation of miR-181a [119] and miR-151 [120] in the dorsal Hip induces long-term hippocampal memory through down-regulation of hippocampal protein levels of protein kinase AMP-activated catalytic subunit alpha 1 (PRKAA1) and anterior pharynx defective 1a (APH1a) receptively following fear conditioning. Moreover, an early decrease in the expression of miR-187-3p was observed in the dorsal Hip following a contextual fear-conditioning paradigm in mice. Interestingly, the downregulation of miR-187-3p is associated with the decrease of stabilin-2 (stab2) protein level [121]. It seems that stress exposure with activating the HPA axis induces modification on neuronal epigenetic regulators to change some hippocampal miRNA-dependent pathways for regulating memory formation. Thus, miRNA expression changes are potentially being used as a novel target for memory dysfunction.

    Stress changes the number and structure of astrocytes

    Astrocytes are the most abundant glial cells in the CNS. They contribute to the blood-brain barrier formation [122], neurotrophin secretion [123], neurotransmitter recycling [124], and regulation of neuronal synaptogenesis [125]. Astrocytes participate in information processing, neuronal plasticity, and LTP via releasing the neuroactive compounds known as gliotransmitters. Astrocytes release D-serine as an endogenous co-agonist of NMDA receptors to enhance the occurrence of NMDA-dependent LTP in nearby excitatory synapses [126]. Additionally, the activation of the astrocytic cAMP enhances the lactate shuttle from astrocytes to neurons, serving as energy for synaptic plasticity and memory formation processes [127]. Interestingly, astrocytes undergo structural plasticity during memory formation. Ostroff and co-workers [128] have shown that astrocytes of the lateral AMY play a critical role in the morphological remodeling of the synapses during implicit memory consolidation. Hence, astrocytes are involved in forming new memories and induction of LTP, and any structural and functional changes will affect the memory formation processes.

    Like neurons and microglia cells, the astrocytes express the GRs and MRs [129]. With the elevation of the stress hormones and extracellular glutamate after HPA axis activation, astrocyte is one of the critical cells that respond to stress’s physiological consequences and are influenced structurally and functionally by stress conditions. Stress induces the hypertrophy of astrocytes and a reduction in gap junction coupling between cells in the Hip and the neocortex. The latter effects reduce functional coupling between hippocampal and neocortex astrocytes which is associated with the attenuation of astrocyte’s capacity to supply neurons with L-lactate. Moreover, the disruption of intracellular networks between the astrocytes is associated with decreased hippocampal LTP and spatial memory impairment [130]. Exposure to chronic stress also reduces the volume fraction of fine astrocytic protrusions and the number of astrocytes in the AMY, which may affect AMY-related memory formation [131]. Astrocytes’ specific elimination of GRs in mice results in impairment of fear memory and aversive memory formation, which suggests that the signaling of the astrocytic GRs regulates stress-induced-aversive memory formation [132]. In general, it can be suggested that astrocytes should be considered a cellular target for therapies for stress-induced memory impairment.

    Stress-induced significant changes in microglia may be involved in memory impairment

    Exposure to stressful conditions and activation of signaling pathways of stress receptors leads to the inhibition of the production of proinflammatory cytokines such as interleukin-1β (IL-1β), IL-18, IL-6, or tumor necrosis factor-α (TNF-α) in the CNS [133]. However, the immune system function disrupts and leads to the overactivation of inflammatory pathways under severely stressful conditions. Microglia cells, the CNS macrophage cells, express the stress hormone receptors, including the glucocorticoid (GC), MR, and norepinephrine (NE) receptors. Therefore, these cells are one kind of the key cells in regulating stress-associated outcomes [134, 135]. Severe and intense stress over-activates the microglia cells and thereby enhances the release of inflammatory cytokines, leading to neuroinflammation. The morphology of microglial cells changes due to stress, and they show fewer branches and an enlarged soma called ameboid microglia [136]. Besides, as a danger signal, stress triggers the formation and activation of microglia inflammasomes. The inflammasomes are intracellular protein structures that involve in the formation of proinflammatory cytokines in cells and play a critical role in innate immunity [137]. The perturbations in neuroimmune function can lead to impaired function of neuronal synapses and deficits of synaptic plasticity that underlie cognitive dysfunction.

    Microglia cells have a pivotal role in memory formation processes by monitoring the neuronal microenvironment, controlling neuronal activity, and neurotransmitter release, maintenance of dendritic spine densities, and mediating the forgetting of remote memories [138]. Because of their role as synaptic sensors that control synaptic development and function, these cells are considered the fourth component of the “quad-partite synapse” in addition to the pre-and postsynaptic terminals and the astrocytes [139]. Previous studies have confirmed that stress could affect memory formation via microglia cells. For example, exposure to chronic restraint stress increases the expression of hippocampal microglia inflammasomes such as Nod-like receptor protein 3 (NLRP3), leading to neuronal injury and memory impairment [140]. The hippocampal expression of NLRP3, IL-1β, and IL-18 was increased, followed by cognitive impairment in socially isolated mice [141]. The microglia cells affect the neurogenesis and maturation of neuronal synaptic in the brain of a mouse model of early life stress. Exposure to stressors dysregulates the microglia function, negatively affecting the neurogenesis and neuronal functions, and may have long-lasting consequences over the lifespan [142].

    Memory impairments associated with stress-related disorders

    A human feels stress from time to time (in everyday life). Among the most potent stressors are psychological and psychosocial stressors, which have unhealthy consequences and adverse effects on the mind and body [143]. Although normal activation of the HPA axis is required for stress response adaption, excessive activation of the HPA axis seems to have detrimental outcomes and is a risk factor for predisposition to several diseases [144, 145]. There is an unanswered question about how stress may affect an individual’s health. Evidence implicated that stress-related neurological responses are closely linked to anxiety disorders such as post-traumatic stress disorders (PTSD) and mood disorders, including major depressive disorder (MDD). Moreover, stress increases the risk of psychiatric disorders such as bipolar and schizophrenia [146, 147]. Besides the negative impact of stress on an individual’s mental illness, long-term exposure to stressors is associated with cognitive dysfunction. It has a prominent role in cognitive decline, leading to amnestic mild cognitive impairment [148] to induce memory problems, particularly in older adults [149].

    Glucocorticoids, as end products of stress-related responses, play a critical role in memory formation via binding to their specific receptors [150152]. Stress hormones interplay between multiple neurotransmitter signaling pathways in the various brain areas. These interactions play a modulatory role in stress circuits, mainly neural circuitry involved in memory performance [152]. It should be noted that the ECS has a modulatory role in glucocorticoid-mediated outcomes to impair cognitive functions during chronic stress [105]. Several observations demonstrated the enhanced endocannabinoid signaling pathways via inhibiting fatty acid amide hydrolase (FAAH), an enzyme responsible for endocannabinoid degradation, to reverse the negative impact on the chronic stress-exposed animals in an object recognition task [153]. Acute stress following memory consolidation led to long-term object recognition memory impairment, while intra-hippocampal microinjection of CB antagonists prevented the memory deficit [37]. The activation of CBs improved memory loss by facilitating the hippocampal LTP [106]. Pre-test intra-dorsal striatal microinjection of WIN 55,212-2, an agonist of CBs, increased memory formation in the passive avoidance task [107]. This study indicated that the central or peripheral blockade of CB1 receptors eliminated the effect of acute stress on passive avoidance memory. These findings suggest that the ECS may be a potential therapeutic target to regulate glucocorticoid-mediated stress memory performance. Furthermore, activating the hippocampal cholinergic system via acetylcholine nicotine receptor signaling pathways is critical for preventing cognitive decline [154]. Keshavarzian et al. [82] reported that pre-test nicotine administration improved acute stress-induced memory impairment. They also showed that the BLA activation of dopamine receptors before memory consolidation reversed the impairing effect of stress on memory retention. The interaction between stress hormones and the dopaminergic system can be suggested to regulate memory formation [155, 156].

    Approximately 30% of stroke patients develop long-term memory impairment within one year of onset [157]. Notably, after AD, vascular dementia is the leading cause of dementia in the world that occurs due to insufficient and impaired blood flow in the brain [158]. On the other hand, stroke is a physiological stressor that activates the HPA axis. Chronic activation of the axis exacerbates stroke outcomes [159]. Clinical studies have shown that chronic stroke patients’ psychological symptoms, such as anxiety or depression, correlate with cognitive dysfunctions [160]. Animal studies have also shown that pre-stroke exposure to psychological stress increases infarct volume and neurological deficits and has detrimental effects on cognitive function [161]. Activation of the HPA axis during the stroke affects the lesion area (mainly the cerebral cortex and striatum) and the contralateral side of the lesion area. Activation of the hippocampal GC receptors in the contralateral side of the ischemic brain leads to hippocampal neuronal damage, which is accomplished by cognitive and psychiatric disturbances that include delayed consequences of stroke [162]. Thus, it can be suggested that targeting stress hormones and their receptors may be a therapeutic approach to ischemia-related cognitive impairments.

    Conclusions

    Stress regulates multiple CNS functions, including memory formation, mood, emotional behaviors, and reward. Evidence suggests that acute and chronic stress change synaptic transmission, neuronal epigenetic modulators, and glial activity in memory-related brain regions. It is important to note that stress develops neurodegenerative disorders with cognitive dysfunctions. Multiple neurotransmission changes happen during memory consolidation, retention, and retrieval under stress. Besides, exposure to severe stressful conditions changes the activity and morphology of astrocytes and microglial cells, which in turn increases the neuroinflammatory cascades and induces neuronal injury. Changing the CNS miRNAs expressions, developing neurodegenerative diseases and cognitive dysfunctions are stress-related responses in humans and laboratory animals. Although various stress mechanisms have been studied in memory impairment following neurodegenerative diseases, some questions remain that are essential to answer for finding appropriate therapies. Therefore, further studies should investigate precise targets of stress signaling pathways that contribute to memory formation. We hope this review will draw more attention to developing treatment strategies for stress-related disorders.

    Abbreviations

    5-HT:

    5-hydroxytryptamine

    5-HT1A:

    5-hydroxytryptamine 1A receptor

    ACTH:

    adrenocorticotropic hormone

    AMY:

    amygdala

    BLA:

    basolateral amygdala

    CBs:

    cannabinoid receptors

    CNS:

    central nervous system

    CRH:

    corticotropin-releasing hormone

    ECS:

    endocannabinoid system

    GABA:

    gamma-aminobutyric acid

    GABAA:

    gamma-aminobutyric acid subtype A

    GABAB:

    gamma-aminobutyric acid subtype B

    GABAergic:

    gamma aminobutyric acidergic

    GRs:

    glucocorticoid receptors

    Hip:

    hippocampus

    HPA:

    hypothalamic-pituitary-adrenocortical

    IL-1β:

    interleukin-1β

    LTP:

    long-term potentiation

    miRNAs:

    microRNAs

    mRNA:

    messenger RNA

    MRs:

    mineralocorticoid

    NMDA:

    N-methyl-D-aspartate

    PFC:

    prefrontal cortex

    PVN:

    paraventricular nucleus

    SAM:

    sympathetic-adrenomedullary

    Declarations

    Author contributions

    AR: Writing—original draft, Designing the figures, Writing—review & editing. MS: Writing—original draft, Designing the figures. SH: Writing—original draft.

    Conflicts of interest

    The authors declare that they have no conflicts of interest.

    Ethical approval

    Not applicable.

    Consent to participate

    Not applicable.

    Consent to publication

    Not applicable.

    Availability of data and materials

    Not applicable.

    Funding

    Not applicable.

    Copyright

    © The Author(s) 2022.

    References

    Charney DS. Psychobiological mechanisms of resilience and vulnerability: implications for successful adaptation to extreme stress. Am J Psychiatry. 2004;161:195216. [DOI] [PubMed]
    Nesse RM, Bhatnagar S, Ellis B. Chapter 11 - Evolutionary origins and functions of the stress response system. In: In: Fink G, editor. Stress: concepts, cognition, emotion, and behavior. San Diego: Academic Press; 2016. pp. 95101. [DOI]
    Ulrich-Lai YM, Herman JP. Neural regulation of endocrine and autonomic stress responses. Nat Rev Neurosci. 2009;10:397409. [DOI] [PubMed] [PMC]
    Joëls M, Baram TZ. The neuro-symphony of stress. Nat Rev Neurosci. 2009;10:45966. [DOI] [PubMed] [PMC]
    Kvetnansky R, Sabban EL, Palkovits M. Catecholaminergic systems in stress: structural and molecular genetic approaches. Physiol Rev. 2009;89:535606. [DOI] [PubMed]
    Ayada C, Toru Ü, Korkut Y. The relationship of stress and blood pressure effectors. Hippokratia. 2015;19:99108. [PubMed] [PMC]
    Herman JP, Flak J, Jankord R. Chronic stress plasticity in the hypothalamic paraventricular nucleus. Prog Brain Res. 2008;170:35364. [DOI] [PubMed]
    Jankord R, Herman JP. Limbic regulation of hypothalamo-pituitary-adrenocortical function during acute and chronic stress. Ann N Y Acad Sci. 2008;1148:6473. [DOI] [PubMed] [PMC]
    Chrousos GP, Kino T, Charmandari E. Evaluation of the hypothalamic-pituitary-adrenal axis function in childhood and adolescence. Neuroimmunomodulation. 2009;16:27283. [DOI] [PubMed] [PMC]
    Higo S, Hojo Y, Ishii H, Komatsuzaki Y, Ooishi Y, Murakami G, et al. Endogenous synthesis of corticosteroids in the hippocampus. PLoS ONE. 2011;6:e21631. [DOI] [PubMed] [PMC]
    Zhe D, Fang H, Yuxiu S. Expressions of hippocampal mineralocorticoid receptor (MR) and glucocorticoid receptor (GR) in the single-prolonged stress-rats. Acta Histochem Cytochem. 2008;41:8995. [DOI] [PubMed] [PMC]
    Eagle AL, Knox D, Roberts MM, Mulo K, Liberzon I, Galloway MP, et al. Single prolonged stress enhances hippocampal glucocorticoid receptor and phosphorylated protein kinase B levels. Neurosci Res. 2013;75:1307. [DOI] [PubMed] [PMC]
    Herman JP, McKlveen JM, Solomon MB, Carvalho-Netto E, Myers B. Neural regulation of the stress response: glucocorticoid feedback mechanisms. Braz J Med Biol Res. 2012;45:2928. [DOI] [PubMed] [PMC]
    Uchoa ET, Aguilera G, Herman JP, Fiedler JL, Deak T, de Sousa MB. Novel aspects of hypothalamic- pituitary-adrenal axis regulation and glucocorticoid actions. J Neuroendocrinol. 2014;26:55772. [DOI] [PubMed] [PMC]
    Herman JP, McKlveen JM, Ghosal S, Kopp B, Wulsin A, Makinson R, et al. Regulation of the hypothalamic-pituitary-adrenocortical stress response. Compr Physiol. 2016;6:60321. [DOI] [PubMed] [PMC]
    Ziegler DR, Herman JP. Neurocircuitry of stress integration: anatomical pathways regulating the hypothalamo-pituitary-adrenocortical axis of the Rat1. Integr Comp Biol. 2002;42:54151. [DOI] [PubMed]
    Berardelli R, Karamouzis I, Marinazzo E, Prats E, Picu A, Giordano R, et al. Effect of acute and prolonged mineralocorticoid receptor blockade on spontaneous and stimulated hypothalamic-pituitary-adrenal axis in humans. Eur J Endocrinol. 2010;162:106774. [DOI] [PubMed]
    Herman JP, Nawreen N, Smail MA, Cotella EM. Brain mechanisms of HPA axis regulation: neurocircuitry and feedback in context Richard Kvetnansky lecture. Stress. 2020;23:61732. [DOI] [PubMed] [PMC]
    MacLullich AM, Ferguson KJ, Wardlaw JM, Starr JM, Deary IJ, Seckl JR. Smaller left anterior cingulate cortex volumes are associated with impaired hypothalamic-pituitary-adrenal axis regulation in healthy elderly men. J Clin Endocrinol Metab. 2006;91:15914. [DOI] [PubMed]
    Kiem SA, Andrade KC, Spoormaker VI, Holsboer F, Czisch M, Sämann PG. Resting state functional MRI connectivity predicts hypothalamus-pituitary-axis status in healthy males. Psychoneuroendocrinology. 2013;38:133848. [DOI] [PubMed]
    Dunn JD, Orr SE. Differential plasma corticosterone responses to hippocampal stimulation. Exp Brain Res. 1984;54:16. [DOI] [PubMed]
    Herman JP, Figueiredo H, Mueller NK, Ulrich-Lai Y, Ostrander MM, Choi DC, et al. Central mechanisms of stress integration: hierarchical circuitry controlling hypothalamo-pituitary-adrenocortical responsiveness. Front Neuroendocrinol. 2003;24:15180. [DOI] [PubMed]
    Smith SM, Vale WW. The role of the hypothalamic-pituitary-adrenal axis in neuroendocrine responses to stress. Dialogues Clin Neurosci. 2006;8:38395. [DOI] [PubMed] [PMC]
    Ressler KJ. Amygdala activity, fear, and anxiety: modulation by stress. Biol Psychiatry. 2010;67:11179. [DOI] [PubMed] [PMC]
    Koning A, Buurstede JC, van Weert L, Meijer OC. Glucocorticoid and mineralocorticoid receptors in the brain: a transcriptional perspective. J Endocr Soc. 2019;3:191730. [DOI] [PubMed] [PMC]
    Joëls M, Sarabdjitsingh RA, Karst H. Unraveling the time domains of corticosteroid hormone influences on brain activity: rapid, slow, and chronic modes. Pharmacol Rev. 2012;64:90138. [DOI] [PubMed]
    Romeo RD. Hypothalamic nuclear glucocorticoid receptors: acute stress and rapid actions. Endocrinology. 2015;156:27478. [DOI] [PubMed]
    McKlveen JM, Myers B, Flak JN, Bundzikova J, Solomon MB, Seroogy KB, et al. Role of prefrontal cortex glucocorticoid receptors in stress and emotion. Biol Psychiatry. 2013;74:6729. [DOI] [PubMed] [PMC]
    Han F, Ding J, Shi Y. Expression of amygdala mineralocorticoid receptor and glucocorticoid receptor in the single-prolonged stress rats. BMC Neurosci. 2014;15:77. [DOI] [PubMed] [PMC]
    Gądek-Michalska A, Spyrka J, Rachwalska P, Tadeusz J, Bugajski J. Influence of chronic stress on brain corticosteroid receptors and HPA axis activity. Pharmacol Rep. 2013;65:116375. [DOI] [PubMed]
    Mayford M, Siegelbaum SA, Kandel ER. Synapses and memory storage. Cold Spring Harb Perspect Biol. 2012;4:a005751. [DOI] [PubMed] [PMC]
    Sardari M, Rezayof A, Khodagholi F, Zarrindast MR. Basolateral amygdala GABA-A receptors mediate stress-induced memory retrieval impairment in rats. Int J Neuropsychopharmacol. 2014;17:60312. [DOI] [PubMed]
    Joëls M. Corticosteroid effects in the brain: U-shape it. Trends Pharmacol Sci. 2006;27:24450. [DOI] [PubMed]
    Sardari M, Rezayof A, Zarrindast MR. 5-HT1A receptor blockade targeting the basolateral amygdala improved stress-induced impairment of memory consolidation and retrieval in rats. Neuroscience. 2015;300:60918. [DOI] [PubMed]
    Schwabe L, Joëls M, Roozendaal B, Wolf OT, Oitzl MS. Stress effects on memory: an update and integration. Neurosci Biobehav Rev. 2012;36:17409. [DOI] [PubMed]
    Wolf OT, Atsak P, de Quervain DJ, Roozendaal B, Wingenfeld K. Stress and memory: a selective review on recent developments in the understanding of stress hormone effects on memory and their clinical relevance. J Neuroendocrinol. 2016;28. [DOI] [PubMed]
    Busquets-Garcia A, Gomis-González M, Srivastava RK, Cutando L, Ortega-Alvaro A, Ruehle S, et al. Peripheral and central CB1 cannabinoid receptors control stress-induced impairment of memory consolidation. Proc Natl Acad Sci U S A. 2016;113:99049. [DOI] [PubMed] [PMC]
    Wirkner J, Weymar M, Löw A, Hamm AO. Effects of pre-encoding stress on brain correlates associated with the long-term memory for emotional scenes. PLoS ONE. 2013;8:e68212. [DOI] [PubMed] [PMC]
    McCullough AM, Ritchey M, Ranganath C, Yonelinas A. Differential effects of stress-induced cortisol responses on recollection and familiarity-based recognition memory. Neurobiol Learn Mem. 2015;123:110. [DOI] [PubMed] [PMC]
    Zoladz PR, Kalchik AE, Hoffman MM, Aufdenkampe RL, Burke HM, Woelke SA, et al. Brief, pre-retrieval stress differentially influences long-term memory depending on sex and corticosteroid response. Brain Cogn. 2014;85:27785. [DOI] [PubMed]
    Mohammadmirzaei N, Rezayof A, Ghasemzadeh Z. Activation of cannabinoid CB1 receptors in the ventral hippocampus improved stress-induced amnesia in rat. Brain Res. 2016;1646:21926. [DOI] [PubMed]
    Sebastian V, Estil JB, Chen D, Schrott LM, Serrano PA. Acute physiological stress promotes clustering of synaptic markers and alters spine morphology in the hippocampus. PLoS ONE. 2013;8:e79077. [DOI] [PubMed] [PMC]
    Christoffel DJ, Golden SA, Russo SJ. Structural and synaptic plasticity in stress-related disorders. Rev Neurosci. 2011;22:53549. [DOI] [PubMed] [PMC]
    Chen Y, Dubé CM, Rice CJ, Baram TZ. Rapid loss of dendritic spines after stress involves derangement of spine dynamics by corticotropin-releasing hormone. J Neurosci. 2008;28:290311. [DOI] [PubMed] [PMC]
    Chen Y, Rex CS, Rice CJ, Dubé CM, Gall CM, Lynch G, et al. Correlated memory defects and hippocampal dendritic spine loss after acute stress involve corticotropin-releasing hormone signaling. Proc Natl Acad Sci U S A. 2010;107:131238. [DOI] [PubMed] [PMC]
    Andres AL, Regev L, Phi L, Seese RR, Chen Y, Gall CM, et al. NMDA receptor activation and calpain contribute to disruption of dendritic spines by the stress neuropeptide CRH. J Neurosci. 2013;33:1694560. [DOI] [PubMed] [PMC]
    Yuen EY, Liu W, Karatsoreos IN, Feng J, McEwen BS, Yan Z. Acute stress enhances glutamatergic transmission in prefrontal cortex and facilitates working memory. Proc Natl Acad Sci U S A. 2009;106:140759. [DOI] [PubMed] [PMC]
    Yuen EY, Liu W, Karatsoreos IN, Ren Y, Feng J, McEwen BS, et al. Mechanisms for acute stress-induced enhancement of glutamatergic transmission and working memory. Mol Psychiatry. 2011;16:15670. [DOI] [PubMed] [PMC]
    Wheelan N, Kenyon CJ, Harris AP, Cairns C, Al Dujaili E, Seckl JR, et al. Midlife stress alters memory and mood-related behaviors in old age: role of locally activated glucocorticoids. Psychoneuroendocrinology. 2018;89:1322. [DOI] [PubMed] [PMC]
    Tripathi SJ, Chakraborty S, Srikumar BN, Raju TR, Shankaranarayana Rao BS. Inactivation of basolateral amygdala prevents stress-induced astroglial loss in the prefrontal cortex. Mol Neurobiol. 2019;56:35066. [DOI] [PubMed]
    Izquierdo I. Effect of beta-endorphin and naloxone on acquisition, memory, and retrieval of shuttle avoidance and habituation learning in rats. Psychopharmacology (Berl). 1980;69:1115. [DOI] [PubMed]
    Shors TJ, Elkabes S, Selcher JC, Black IB. Stress persistently increases NMDA receptor-mediated binding of [3H]PDBu (a marker for protein kinase C) in the amygdala, and re-exposure to the stressful context reactivates the increase. Brain Res. 1997;750:293300. [DOI] [PubMed]
    Bartsch JC, von Cramon M, Gruber D, Heinemann U, Behr J. Stress-induced enhanced long-term potentiation and reduced threshold for N-methyl-D-aspartate receptor- and β-adrenergic receptor-mediated synaptic plasticity in rodent ventral subiculum. Front Mol Neurosci. 2021;14:658465. [DOI] [PubMed] [PMC]
    Popoli M, Yan Z, McEwen BS, Sanacora G. The stressed synapse: the impact of stress and glucocorticoids on glutamate transmission. Nat Rev Neurosci. 2011;13:2237. [DOI] [PubMed] [PMC]
    Maggio N, Segal M. Differential corticosteroid modulation of inhibitory synaptic currents in the dorsal and ventral hippocampus. J Neurosci. 2009;29:285766. [DOI] [PubMed] [PMC]
    Maggio N, Segal M. Persistent changes in ability to express long-term potentiation/depression in the rat hippocampus after juvenile/adult stress. Biol Psychiatry. 2011;69:74853. [DOI] [PubMed]
    Grigoryan G, Ardi Z, Albrecht A, Richter-Levin G, Segal M. Juvenile stress alters LTP in ventral hippocampal slices: involvement of noradrenergic mechanisms. Behav Brain Res. 2015;278:55962. [DOI] [PubMed]
    Joffe ME, Santiago CI, Oliver KH, Maksymetz J, Harris NA, Engers JL, et al. mGlu2 and mGlu3 negative allosteric modulators divergently enhance thalamocortical transmission and exert rapid antidepressant-like effects. Neuron. 2020;105:4659.E3. [DOI] [PubMed] [PMC]
    Ryan S, Li C, Menigoz A, Hazra R, Dabrowska J, Ehrlich D, et al. Repeated shock stress facilitates basolateral amygdala synaptic plasticity through decreased cAMP-specific phosphodiesterase type IV (PDE4) expression. Brain Struct Funct. 2018;223:173145. [DOI] [PubMed] [PMC]
    Shavit Stein E, Itsekson Hayosh Z, Vlachos A, Maggio N. Stress and corticosteroids modulate muscarinic long term potentiation (mLTP) in the hippocampus. Front Cell Neurosci. 2017;11:299. [DOI] [PubMed] [PMC]
    Loprinzi PD, Frith E. Protective and therapeutic effects of exercise on stress-induced memory impairment. J Physiol Sci. 2019;69:112. [DOI] [PubMed]
    Mora F, Segovia G, Del Arco A, de Blas M, Garrido P. Stress, neurotransmitters, corticosterone and body-brain integration. Brain Res. 2012;1476:7185. [DOI] [PubMed]
    Savage K, Firth J, Stough C, Sarris J. GABA-modulating phytomedicines for anxiety: a systematic review of preclinical and clinical evidence. Phytother Res. 2018;32:318. [DOI] [PubMed]
    Sarawagi A, Soni ND, Patel AB. Glutamate and GABA homeostasis and neurometabolism in major depressive disorder. Front Psychiatry. 2021;12:637863. [DOI] [PubMed] [PMC]
    Kotak VC, Mirallave A, Mowery TM, Sanes DH. GABAergic inhibition gates excitatory LTP in perirhinal cortex. Hippocampus. 2017;27:121723. [DOI] [PubMed] [PMC]
    Sanchez-Vives MV, Barbero-Castillo A, Perez-Zabalza M, Reig R. GABAB receptors: modulation of thalamocortical dynamics and synaptic plasticity. Neuroscience. 2021;456:13142. [DOI] [PubMed]
    de Groote L, Linthorst AC. Exposure to novelty and forced swimming evoke stressor-dependent changes in extracellular GABA in the rat hippocampus. Neuroscience. 2007;148:794805. [DOI] [PubMed]
    Reznikov LR, Reagan LP, Fadel JR. Effects of acute and repeated restraint stress on GABA efflux in the rat basolateral and central amygdala. Brain Res. 2009;1256:618. [DOI] [PubMed]
    Herman JP, Mueller NK, Figueiredo H. Role of GABA and glutamate circuitry in hypothalamo-pituitary-adrenocortical stress integration. Ann N Y Acad Sci. 2004;1018:3545. [DOI] [PubMed]
    Shang Y, Chen R, Li F, Zhang H, Wang H, Zhang T. Prenatal stress impairs memory function in the early development of male-offspring associated with the gaba function. Physiol Behav. 2021;228:113184. [DOI] [PubMed]
    Beaulieu JM, Gainetdinov RR. The physiology, signaling, and pharmacology of dopamine receptors. Pharmacol Rev. 2011;63:182217. [DOI] [PubMed]
    Hall H, Sedvall G, Magnusson O, Kopp J, Halldin C, Farde L. Distribution of D1- and D2-dopamine receptors, and dopamine and its metabolites in the human brain. Neuropsychopharmacology. 1994;11:24556. [DOI] [PubMed]
    Rezazadeh M, Ahmadifar M, Manesh MA. The study of effect of amphetamine on passive avoidance learning in Wistar male rats. Adv Appl Physiol. 2018;3:17. [DOI]
    McNamara CG, Tejero-Cantero Á, Trouche S, Campo-Urriza N, Dupret D. Dopaminergic neurons promote hippocampal reactivation and spatial memory persistence. Nat Neurosci. 2014;17:165860. [DOI] [PubMed] [PMC]
    Baudonnat M, Huber A, David V, Walton ME. Heads for learning, tails for memory: reward, reinforcement and a role of dopamine in determining behavioral relevance across multiple timescales. Front Neurosci. 2013;7:175. [DOI] [PubMed] [PMC]
    Baik JH. Stress and the dopaminergic reward system. Exp Mol Med. 2020;52:187990. [DOI] [PubMed] [PMC]
    Saal D, Dong Y, Bonci A, Malenka RC. Drugs of abuse and stress trigger a common synaptic adaptation in dopamine neurons. Neuron. 2003;37:57782. Erratum in: Neuron. 2003;38:359. [DOI] [PubMed]
    Sanchez CJ, Bailie TM, Wu WR, Li N, Sorg BA. Manipulation of dopamine d1-like receptor activation in the rat medial prefrontal cortex alters stress- and cocaine-induced reinstatement of conditioned place preference behavior. Neuroscience. 2003;119:497505. [DOI] [PubMed]
    Bahari Z, Meftahi GH, Meftahi MA. Dopamine effects on stress-induced working memory deficits. Behav Pharmacol. 2018;29:58491. [DOI] [PubMed]
    Gamo NJ, Lur G, Higley MJ, Wang M, Paspalas CD, Vijayraghavan S, et al. Stress impairs prefrontal cortical function via D1 dopamine receptor interactions with hyperpolarization-activated cyclic nucleotide-gated channels. Biol Psychiatry. 2015;78:86070. [DOI] [PubMed] [PMC]
    Pakdel R, Rashidy-Pour A. Glucocorticoid-induced impairment of long-term memory retrieval in rats: an interaction with dopamine D2 receptors. Neurobiol Learn Mem. 2006;85:3006. [DOI] [PubMed]
    Keshavarzian E, Ghasemzadeh Z, Rezayof A. The basolateral amygdala dopaminergic system contributes to the improving effect of nicotine on stress-induced memory impairment in rats. Prog Neuropsychopharmacol Biol Psychiatry. 2018;86:305. [DOI] [PubMed]
    Nichols DE, Nichols CD. Serotonin receptors. Chem Rev. 2008;108:161441. [DOI] [PubMed]
    Fink KB, Göthert M. 5-HT receptor regulation of neurotransmitter release. Pharmacol Rev. 2007;59:360417. Erratum in: Pharmacol Rev. 2008;60:142. [DOI] [PubMed]
    Cavallaro S. Genomic analysis of serotonin receptors in learning and memory. Behav Brain Res. 2008;195:26. [DOI] [PubMed]
    Roth MK, Bingham B, Shah A, Joshi A, Frazer A, Strong R, et al. Effects of chronic plus acute prolonged stress on measures of coping style, anxiety, and evoked HPA-axis reactivity. Neuropharmacology. 2012;63:111826. [DOI] [PubMed] [PMC]
    Yang L, Zhao Y, Wang Y, Liu L, Zhang X, Li B, et al. The effects of psychological stress on depression. Curr Neuropharmacol. 2015;13:494504. [DOI] [PubMed] [PMC]
    Kim YR, Park BK, Kim YH, Shim I, Kang IC, Lee MY. Antidepressant effect of Fraxinus rhynchophylla hance extract in a mouse model of chronic stress-induced depression. Biomed Res Int. 2018;2018:8249563. Erratum in: Biomed Res Int. 2019;2019:4672059. [DOI] [PubMed] [PMC]
    Léa Blondelle KD, Simplice FH, Hervé Hervé NA, Eglantine KW, Roland RN, Jorelle Linda DK, et al. Antidepressant, anti-amnesic and vasoprotective effect of Bombax costatum Pellegr. & Vuillet aqueous stem bark extract on chronic mild unpredictable stress induced in rat. J Ethnopharmacol. 2022;293:115315. [DOI] [PubMed]
    Natarajan R, Forrester L, Chiaia NL, Yamamoto BK. Chronic-stress-induced behavioral changes associated with subregion-selective serotonin cell death in the dorsal raphe. J Neurosci. 2017;37:621423. [DOI] [PubMed] [PMC]
    Baratta MV, Kodandaramaiah SB, Monahan PE, Yao J, Weber MD, Lin PA, et al. Stress enables reinforcement-elicited serotonergic consolidation of fear memory. Biol Psychiatry. 2016;79:81422. [DOI] [PubMed] [PMC]
    Ohmura Y, Yoshida T, Konno K, Minami M, Watanabe M, Yoshioka M. Serotonin 5-HT7 receptor in the ventral hippocampus modulates the retrieval of fear memory and stress-induced defecation. Int J Neuropsychopharmacol. 2016;19:pyv131. [DOI] [PubMed] [PMC]
    Zou S, Kumar U. Cannabinoid receptors and the endocannabinoid system: signaling and function in the central nervous system. Int J Mol Sci. 2018;19:833. [DOI] [PubMed] [PMC]
    Ohno-Shosaku T, Kano M. Endocannabinoid-mediated retrograde modulation of synaptic transmission. Curr Opin Neurobiol. 2014;29:18. [DOI] [PubMed]
    Da S, Takahashi RN. SR 141716A prevents delta 9-tetrahydrocannabinol-induced spatial learning deficit in a Morris-type water maze in mice. Prog Neuropsychopharmacol Biol Psychiatry. 2002;26:3215. [DOI] [PubMed]
    Wise LE, Thorpe AJ, Lichtman AH. Hippocampal CB1 receptors mediate the memory impairing effects of Δ9-tetrahydrocannabinol. Neuropsychopharmacology. 2009;34:207280. [DOI] [PubMed] [PMC]
    Ranganathan M, D’Souza DC. The acute effects of cannabinoids on memory in humans: a review. Psychopharmacology (Berl). 2006;188:42544. [DOI] [PubMed]
    Pamplona FA, Prediger RD, Pandolfo P, Takahashi RN. The cannabinoid receptor agonist WIN 55,212-2 facilitates the extinction of contextual fear memory and spatial memory in rats. Psychopharmacology (Berl). 2006;188:6419. [DOI] [PubMed]
    Akirav I. The role of cannabinoids in modulating emotional and non-emotional memory processes in the hippocampus. Front Behav Neurosci. 2011;5:34. [DOI] [PubMed] [PMC]
    Segev A, Akirav I. Differential effects of cannabinoid receptor agonist on social discrimination and contextual fear in amygdala and hippocampus. Learn Mem. 2011;18:2549. [DOI] [PubMed]
    Di Marzo V, Stella N, Zimmer A. Endocannabinoid signalling and the deteriorating brain. Nat Rev Neurosci. 2015;16:3042. [DOI] [PubMed] [PMC]
    Aso E, Palomer E, Juvés S, Maldonado R, Muñoz FJ, Ferrer I. CB1 agonist ACEA protects neurons and reduces the cognitive impairment of AβPP/PS1 mice. J Alzheimers Dis. 2012;30:43959. Erratum in: J Alzheimers Dis. 2012;31:679–80. [DOI] [PubMed]
    Haghani M, Shabani M, Javan M, Motamedi F, Janahmadi M. CB1 cannabinoid receptor activation rescues amyloid β-induced alterations in behaviour and intrinsic electrophysiological properties of rat hippocampal CA1 pyramidal neurones. Cell Physiol Biochem. 2012;29:391406. [DOI] [PubMed]
    Rubino T, Realini N, Castiglioni C, Guidali C, Viganó D, Marras E, et al. Role in anxiety behavior of the endocannabinoid system in the prefrontal cortex. Cereb Cortex. 2008;18:1292301. [DOI] [PubMed]
    Scarante FF, Vila-Verde C, Detoni VL, Ferreira-Junior NC, Guimarães FS, Campos AC. Cannabinoid modulation of the stressed hippocampus. Front Mol Neurosci. 2017;10:411. [DOI] [PubMed] [PMC]
    Shoshan N, Segev A, Abush H, Mizrachi Zer-Aviv T, Akirav I. Cannabinoids prevent the differential long-term effects of exposure to severe stress on hippocampal- and amygdala-dependent memory and plasticity. Hippocampus. 2017;27:1093109. [DOI] [PubMed]
    Siller-Pérez C, Fuentes-Ibañez A, Sotelo-Barrera EL, Serafín N, Prado-Alcalá RA, Campolongo P, et al. Glucocorticoid interactions with the dorsal striatal endocannabinoid system in regulating inhibitory avoidance memory. Psychoneuroendocrinology. 2019;99:97103. [DOI] [PubMed]
    Bhaskaran M, Mohan M. MicroRNAs: history, biogenesis, and their evolving role in animal development and disease. Vet Pathol. 2014;51:75974. [DOI] [PubMed] [PMC]
    Statello L, Guo CJ, Chen LL, Huarte M. Gene regulation by long non-coding RNAs and its biological functions. Nat Rev Mol Cell Biol. 2021;22:96118. Erratum in: Nat Rev Mol Cell Biol. 2021;22:159. [DOI] [PubMed] [PMC]
    Guo H, Ingolia NT, Weissman JS, Bartel DP. Mammalian microRNAs predominantly act to decrease target mRNA levels. Nature. 2010;466:83540. [DOI] [PubMed] [PMC]
    Volk N, Pape JC, Engel M, Zannas AS, Cattane N, Cattaneo A, et al. Amygdalar microRNA-15a is essential for coping with chronic stress. Cell Reports. 2016;17:188291. [DOI] [PubMed] [PMC]
    Haramati S, Navon I, Issler O, Ezra-Nevo G, Gil S, Zwang R, et al. MicroRNA as repressors of stress-induced anxiety: the case of amygdalar miR-34. J Neurosci. 2011;31:14191203. [DOI] [PubMed] [PMC]
    Li C, Liu Y, Liu D, Jiang H, Pan F. Dynamic alterations of miR-34c expression in the hypothalamus of male rats after early adolescent traumatic stress. Neural Plasticity. 2016;2016:5249893. [DOI] [PubMed] [PMC]
    Kao YC, Wang IF, Tsai KJ. miRNA-34c overexpression causes dendritic loss and memory decline. Int J Mol Sci. 2018;19:2323. [DOI] [PubMed] [PMC]
    Andolina D, Di Segni M, Ventura R. MiRNA-34 and stress response. Oncotarget. 2017;8:56589. [DOI] [PubMed] [PMC]
    Liu Y, Liu D, Xu J, Jiang H, Pan F. Early adolescent stress-induced changes in prefrontal cortex miRNA-135a and hippocampal miRNA-16 in male rats. Dev Psychobiol. 2017;59:95869. [DOI] [PubMed]
    Issler O, Haramati S, Paul ED, Maeno H, Navon I, Zwang R, et al. MicroRNA 135 is essential for chronic stress resiliency, antidepressant efficacy, and intact serotonergic activity. Neuron. 2014;83:34460. [DOI] [PubMed]
    Mannironi C, Camon J, De Vito F, Biundo A, De Stefano ME, Persiconi I, et al. Acute stress alters amygdala microRNA miR-135a and miR-124 expression: inferences for corticosteroid dependent stress response. PLoS ONE. 2013;8:e73385. [DOI] [PubMed] [PMC]
    Zhang SF, Chen JC, Zhang J, Xu JG. miR-181a involves in the hippocampus-dependent memory formation via targeting PRKAA1. Sci Rep. 2017;7:8480. [DOI] [PubMed] [PMC]
    Xu XF, Wang YC, Zong L, Wang XL. miR-151-5p modulates APH1a expression to participate in contextual fear memory formation. RNA Biol. 2019;16:28294. Erratum in: RNA Biol. 2020;17:1827. [DOI] [PubMed] [PMC]
    Zhao C, Zhou B, Cao J, Zhang Y, Li W, Wang M, et al. miR-187-3p participates in contextual fear memory formation through modulating SATB2 expression in the hippocampus. Neuroreport. 2020;31:90917. [DOI] [PubMed]
    Cabezas R, Avila M, Gonzalez J, El-Bachá RS, Báez E, García-Segura LM, et al. Astrocytic modulation of blood brain barrier: perspectives on Parkinson’s disease. Front Cell Neurosci. 2014;8:211. [DOI] [PubMed] [PMC]
    Pöyhönen S, Er S, Domanskyi A, Airavaara M. Effects of neurotrophic factors in glial cells in the central nervous system: expression and properties in neurodegeneration and injury. Front Physiol. 2019;10:486. [DOI] [PubMed] [PMC]
    Andersen JV, Markussen KH, Jakobsen E, Schousboe A, Waagepetersen HS, Rosenberg PA, et al. Glutamate metabolism and recycling at the excitatory synapse in health and neurodegeneration. Neuropharmacology. 2021;196:108719. [DOI] [PubMed]
    Shan L, Zhang T, Fan K, Cai W, Liu H. Astrocyte-neuron signaling in synaptogenesis. Front Cell Dev Biol. 2021;9:680301. [DOI] [PubMed] [PMC]
    Henneberger C, Papouin T, Oliet SH, Rusakov DA. Long-term potentiation depends on release of D-serine from astrocytes. Nature. 2010;463:2326. [DOI] [PubMed] [PMC]
    Zhou Z, Okamoto K, Onodera J, Hiragi T, Andoh M, Ikawa M, et al. Astrocytic cAMP modulates memory via synaptic plasticity. Proc Natl Acad Sci U S A. 2021;118:e2016584118. [DOI] [PubMed] [PMC]
    Ostroff LE, Manzur MK, Cain CK, Ledoux JE. Synapses lacking astrocyte appear in the amygdala during consolidation of Pavlovian threat conditioning. J Comp Neurol. 2014;522:215263. [DOI] [PubMed] [PMC]
    Wang Q, Verweij EW, Krugers HJ, Joels M, Swaab DF, Lucassen PJ. Distribution of the glucocorticoid receptor in the human amygdala; changes in mood disorder patients. Brain Struct Funct. 2014;219:161526. [DOI] [PubMed]
    Hösli L, Binini N, Ferrari KD, Thieren L, Looser ZJ, Zuend M, et al. Decoupling astrocytes in adult mice impairs synaptic plasticity and spatial learning. Cell Rep. 2022;38:110484. [DOI] [PubMed]
    Naskar S, Chattarji S. Stress elicits contrasting effects on the structure and number of astrocytes in the amygdala versus hippocampus. eNeuro. 2019;6:ENEURO.0338-18.2019. [DOI] [PubMed] [PMC]
    Tertil M, Skupio U, Barut J, Dubovyk V, Wawrzczak-Bargiela A, Soltys Z, et al. Glucocorticoid receptor signaling in astrocytes is required for aversive memory formation. Transl Psychiatry. 2018;8:255. [DOI] [PubMed] [PMC]
    Liberman AC, Budziñski ML, Sokn C, Gobbini RP, Steininger A, Arzt E. Regulatory and mechanistic actions of glucocorticoids on T and inflammatory cells. Front Endocrinol (Lausanne). 2018;9:235. [DOI] [PubMed] [PMC]
    Tanaka J, Fujita H, Matsuda S, Toku K, Sakanaka M, Maeda N. Glucocorticoid- and mineralocorticoid receptors in microglial cells: the two receptors mediate differential effects of corticosteroids. Glia. 1997;20:2337. [DOI]
    Mori K, Ozaki E, Zhang B, Yang L, Yokoyama A, Takeda I, et al. Effects of norepinephrine on rat cultured microglial cells that express α1, α2, β1 and β2 adrenergic receptors. Neuropharmacology. 2002;43:102634. [DOI] [PubMed]
    Wohleb ES, Delpech JC. Dynamic cross-talk between microglia and peripheral monocytes underlies stress-induced neuroinflammation and behavioral consequences. Prog Neuropsychopharmacol Biol Psychiatry. 2017;79:408. [DOI] [PubMed]
    Dong Y, Li S, Lu Y, Li X, Liao Y, Peng Z, et al. Stress-induced NLRP3 inflammasome activation negatively regulates fear memory in mice. J Neuroinflammation. 2020;17:205. [DOI] [PubMed] [PMC]
    Gentry NW, McMahon T, Yamazaki M, Webb J, Arnold TD, Rosi S, et al. Microglia are involved in the protection of memories formed during sleep deprivation. Neurobiol Sleep Circadian Rhythms. 2022;12:100073. [DOI] [PubMed] [PMC]
    Schafer DP, Lehrman EK, Stevens B. The “quad-partite” synapse: microglia-synapse interactions in the developing and mature CNS. Glia. 2013;61:2436. [DOI] [PubMed] [PMC]
    Feng X, Zhao Y, Yang T, Song M, Wang C, Yao Y, et al. Glucocorticoid-driven NLRP3 inflammasome activation in hippocampal microglia mediates chronic stress-induced depressive-like behaviors. Front Mol Neurosci. 2019;12:210. [DOI] [PubMed] [PMC]
    Niu L, Luo SS, Xu Y, Wang Z, Luo D, Yang H, et al. The critical role of the hippocampal NLRP3 inflammasome in social isolation-induced cognitive impairment in male mice. Neurobiol Learn Mem. 2020;175:107301. [DOI] [PubMed]
    Delpech JC, Wei L, Hao J, Yu X, Madore C, Butovsky O, et al. Early life stress perturbs the maturation of microglia in the developing hippocampus. Brain Behav Immun. 2016;57:7993. [DOI] [PubMed] [PMC]
    Kogler L, Müller VI, Chang A, Eickhoff SB, Fox PT, Gur RC, et al. Psychosocial versus physiological stress - meta-analyses on deactivations and activations of the neural correlates of stress reactions. Neuroimage. 2015;119:23551. [DOI] [PubMed] [PMC]
    Stephens MA, Wand G. Stress and the HPA axis: role of glucocorticoids in alcohol dependence. Alcohol Res. 2012;34:46883. [PubMed] [PMC]
    Bermejo JL, Valldecabres R, Villarrasa-Sapiña I, Monfort-Torres G, Marco-Ahulló A, Ribeiro Do Couto B. Increased cortisol levels caused by acute resistance physical exercise impair memory and learning ability. PeerJ. 2022;10:e13000. [DOI] [PubMed] [PMC]
    Davis MT, Holmes SE, Pietrzak RH, Esterlis I. Neurobiology of chronic stress-related psychiatric disorders: evidence from molecular imaging studies. Chronic Stress. 2017;1. [DOI] [PubMed] [PMC]
    Musazzi L, Tornese P, Sala N, Popoli M. What acute stress protocols can tell us about PTSD and stress-related neuropsychiatric disorders. Front Pharmacol. 2018;9:758. [DOI] [PubMed] [PMC]
    Peavy GM, Jacobson MW, Salmon DP, Gamst AC, Patterson TL, Goldman S, et al. The influence of chronic stress on dementia-related diagnostic change in older adults. Alzheimer Dis Assoc Disord. 2012;26:2606. [DOI] [PubMed] [PMC]
    Katz MJ, Derby CA, Wang C, Sliwinski MJ, Ezzati A, Zimmerman ME, et al. Influence of perceived stress on incident amnestic mild cognitive impairment: results from the Einstein aging study. Alzheimer Dis Assoc Disord. 2016;30:938. [DOI] [PubMed] [PMC]
    Finsterwald C, Alberini CM. Stress and glucocorticoid receptor-dependent mechanisms in long-term memory: from adaptive responses to psychopathologies. Neurobiol Learn Mem. 2014;112:1729. [DOI] [PubMed] [PMC]
    Rashid H, Ahmed T. Gender dependent contribution of muscarinic receptors in memory retrieval under sub-chronic stress. Neurosci Lett. 2018;681:611. [DOI] [PubMed]
    Bahtiyar S, Gulmez Karaca K, Henckens MJAG, Roozendaal B. Norepinephrine and glucocorticoid effects on the brain mechanisms underlying memory accuracy and generalization. Mol Cell Neurosci. 2020;108:103537. [DOI] [PubMed]
    Griebel G, Stemmelin J, Lopez-Grancha M, Fauchey V, Slowinski F, Pichat P, et al. The selective reversible FAAH inhibitor, SSR411298, restores the development of maladaptive behaviors to acute and chronic stress in rodents. Sci Rep. 2018;8:2416. [DOI] [PubMed] [PMC]
    Maurer SV, Williams CL. The cholinergic system modulates memory and hippocampal plasticity via its interactions with non-neuronal cells. Front Immunol. 2017;8:1489. [DOI] [PubMed] [PMC]
    Rasheed N, Alghasham A. Central dopaminergic system and its implications in stress-mediated neurological disorders and gastric ulcers: short review. Adv Pharmacol Sci. 2012;2012:182671. [DOI] [PubMed] [PMC]
    Alghasham A, Rasheed Z. Therapeutic targets for rheumatoid arthritis: progress and promises. Autoimmunity. 2014;47:7794. [DOI] [PubMed]
    Cullen B, O’Neill B, Evans JJ, Coen RF, Lawlor BA. A review of screening tests for cognitive impairment. J Neurol Neurosurg Psychiatry. 2007;78:7909. [DOI] [PubMed] [PMC]
    Iadecola C. The pathobiology of vascular dementia. Neuron. 2013;80:84466. [DOI] [PubMed] [PMC]
    Olsson T, Marklund N, Gustafson Y, Näsman B. Abnormalities at different levels of the hypothalamic- pituitary-adrenocortical axis early after stroke. Stroke. 1992;23:15736. [DOI] [PubMed]
    Lajevardi L, Taghizade G, Parnain Z. The effect of psychological factors on cognitive functions in stroke patients with chronic fatigue. Funct Dis J. 2021;15:18592. [DOI]
    Sugo N, Hurn PD, Morahan MB, Hattori K, Traystman RJ, DeVries AC. Social stress exacerbates focal cerebral ischemia in mice. Stroke. 2002;33:16604. [DOI] [PubMed]
    Gulyaeva NV. Functional neurochemistry of the ventral and dorsal hippocampus: stress, depression, dementia and remote hippocampal damage. Neurochem Res. 2019;44:130622. [DOI] [PubMed]