Defects in cytokine-mediated neuroprotective glial responses to excitotoxic hippocampal injury in senescence-accelerated mouse

https://doi.org/10.1016/j.bbi.2010.08.006Get rights and content

Abstract

Aging is a result of damage accumulation, and understanding of the mechanisms of aging requires exploration of the cellular and molecular systems functioning to control damage. Senescence-accelerated mouse prone 10 (SAMP10) has been established as an inbred strain exhibiting accelerated aging with an earlier onset of cognitive impairment due to neurodegeneration than the senescence-resistant control (SAMR1) strain. We hypothesized that tissue-protective responses of glial cells are impaired in SAMP10 mice. We injected kainic acid (KA) to induce hippocampal injury and studied how cytokines were upregulated on Day 3 using 3-month-old SAMP10 and SAMR1 mice. Following microarray-based screening for upregulated genes, we performed real-time RT-PCR and immunohistochemistry. Results indicated well-orchestrated cytokine-mediated glial interactions in the injured hippocampus of SAMR1 mice, in which microglia-derived interferon (IFN)-γ stimulated astrocytes via IFN-γ receptor and thereby induced expression of CXCL10 and macrophage inflammatory protein (MIP)-1α, and activated microglia produced granulocyte–macrophage colony-stimulating factor (GM-CSF) and osteopontin (OPN). OPN was the most strongly upregulated cytokine. CD44, an OPN receptor, was also strongly upregulated in the neuropil, especially on neurons and astrocytes. KA-induced hippocampal upregulation of these cytokines was strikingly reduced in SAMP10 mice compared to SAMR1 mice. On Day 30 after KA injection, SAMP10 but not SAMR1 mice exhibited hippocampal layer atrophy. Since the OPN–CD44 system is essential for neuroprotection and remodeling, these findings highlight the defects of SAMP10 mice in cytokine-mediated neuroprotective glia–neuron interactions, which may be associated with the mechanism underlying the vulnerability of SAMP10 mice to age-related neurodegeneration.

Research highlights

► The SAMP10 mouse is a model of accelerated brain aging. ► Microglia of SAMP10 mice failed to produce IFN-gamma and osteopontin (OPN) following excitotoxic injury. ► Injured hippocampal tissue of SAMP10 mice failed to upregulate the expression of CD44 (which mediates the neuroprotective effects of OPN signaling). ► Defects in cytokine-mediated glial neuroprotection may underlie the age-related neurodegeneration in SAMP10 mice.

Introduction

Brain function is based on complex intercellular signaling networks composed not only of neurons linked through synapses but also glial syncytia through which Ca2+ waves can pass (Scemes and Giaume, 2006). The microglia are a third type of cell that may participate in this intercellular signaling network (Bessis et al., 2007, Hung et al., 2010). Microglia are the resident immune cells of the central nervous system (CNS) and play major roles in the CNS immune network. Microglia detect normal and pathological levels of activity in the brain with short-term and long-term effects on how signals are transmitted through neuronal networks (Moriguchi et al., 2003). Microglia can be quickly activated to release inflammatory mediators, engage in phagocytosis, and participate in adaptive immunity. Astrocytes as well as microglia are rapidly activated in response to high levels of neuronal activity. Astrocytes strongly communicate with microglia and play a critical role in the activation of microglia. There is a good deal of evidence suggesting that interaction occurs between astrocytes and microglia (Lynch, 2009). Cytokine expression by astrocytes appears to be involved in microglia activation (Graeber and Streit, 2010).

Kainic acid (KA), a glutamate analogue, induces hippocampal injury via its excitotoxic effects and causes proliferation and activation of astroglial and microglial cells (Jorgensen et al., 1993). Pyramidal cells in the cornu ammonis (CA) 3 sector are particularly vulnerable to excitotoxicity in mice and rats treated with systemic, intracerebroventricular, or intrahippocampal injection of KA (Ben-Ari, 1985, Jeon et al., 2008, Oprica et al., 2003, Tauck and Nadler, 1985). When CA3 predominant hippocampal damage is induced by KA injection, CA3 pyramidal cells undergo degeneration and cell death, which result in axonal degeneration of the afferents to the CA1 pyramidal cell dendrites (Jorgensen et al., 1990, Nadler et al., 1980). In CA1, neuronal cell bodies are preserved but most of the afferent terminals are at least transiently lost. Neural degeneration gives rise to a characteristic set of morphological transformations of microglia into bushy cells in the stratum radiatum (s.r.) of CA1 and stellate cells in the s.r. of CA3. Neural degeneration also induces hypertrophic changes in astrocytes. These cellular transformations together with the upregulation of a wide variety of intercellular signaling molecules are known to occur most strongly around 3 days after KA injection (Borges et al., 2004, Kim et al., 2002), although early events mediated by pro-inflammatory cytokines start several hours after KA injection (Oprica et al., 2003).

Much evidence suggests that aging is the result of many forms of damage accumulation. Understanding of the cellular and molecular basis of aging requires exploration of the mechanisms causing damage to accumulate and the systems controlling such damage (Kirkwood, 2005). In this context, the immune system plays critical roles in aging, and aging may be associated with altered immune responses in the brain. Steady-state levels of inflammatory cytokines are increased in the brain with advancing age (Sparkman and Johnson, 2008). Microglia may acquire a reactive phenotype during aging (Dilger and Johnson, 2008). Conversely, microglia may not function efficiently in older animals (Conde and Streit, 2006).

The senescence-accelerated mouse (SAM) is a well-characterized animal model of human aging (Sprott and Austad, 2006) and has been widely used to investigate a variety of disorders associated with advancing age (Takeda et al., 1991, Takeda et al., 1981). SAM consists of prone (SAMP) strains that exhibit an accelerated senescence-prone phenotype that includes shortened life span and systemic age-related degenerative diseases as well as resistant (SAMR) strains that are senescence-resistant and have been used as usual aging controls. The SAMP10 strain has been established as an experimental model for the study of brain aging. Aged SAMP10 mice are characterized by impairments in cognitive behavior associated with cerebral atrophy (Shimada, 1999, Shimada et al., 1994, Shimada et al., 1992, Shimada et al., 1993). Individual neurons in the cerebral cortex of aged SAMP10 mice exhibit decreases in soma size, dendritic retraction, and loss of dendritic spines, resulting in loss of synapses (Shimada et al., 2003, Shimada et al., 2006), consistent with changes that occur in the aging human brain (Dickstein et al., 2007). Age-related progressive neuronal DNA damage, neuronal ubiquitinated inclusions, decreased tissue proteasome activity, and elevation of pro-inflammatory cytokine gene expression have been reported in SAMP10 mice (Kumagai et al., 2007, Shimada et al., 2008, Shimada et al., 2002). These microscopically and biochemically identifiable types of degeneration of neurons begin to appear around the age of 8 months in SAMP10 mice. Other strains of mice with usual aging processes such as SAMR1 and C57BL/6 do not exhibit remarkable age-related neurodegeneration. Our recent study revealed that hippocampal microglia of young SAMP10 mice have degenerated cytoplasmic processes similar to those in aged SAMR1 mice and more frequent pathological changes such as beading and fragmentation of processes (Hasegawa-Ishii et al., 2010). These morphological abnormalities in microglia of SAMP10 mice appear at the age of 3 months and precede the onset of neuronal degeneration.

In the present study, we hypothesized that intercellular signaling networks comprised of neurons, astrocytes, and microglia are impaired in young SAMP10 mice, and that dysfunction in the CNS immune network that has neuroprotective functions against tissue injury underlies the accelerated brain aging observed in this model. We injected KA into SAMP10 and SAMR1 mice and screened for upregulated cytokine-related genes, and identified tissue and cellular sources of upregulated molecules. Findings were compared between SAMP10 and SAMR1 mice to highlight the potential abnormalities in cytokine responses to excitotoxic hippocampal injury in SAMP10 mice.

Section snippets

Animals

Founder pairs of inbred SAMP10 and SAMR1 strains of mice were provided by The Council for SAM Research (Kyoto, Japan) to establish colonies of aging mice at the Institute for Developmental Research. The SAMP10 aging colony (SAMP10/Idr) and the SAMR1 aging colony (SAMR1/Idr) were expanded and maintained by brother–sister mating in our animal house at a temperature of 23 ± 2 °C and humidity of 50 ± 10%. Male SAMP10 and SAMR1 mice were at the age of 3 months intraperitoneally injected with either 40 

KA-induced hippocampal injury

In the hippocampus of mice injected with KA, pyramidal cells of the CA3 sector underwent degeneration and cell death (Fig. 1A–C). In the CA1 sector, cell bodies of pyramidal cells appeared to be relatively preserved (Fig. 2A–C). There was no evidence of inflammatory cell infiltrates such as polymorphonuclear cells or lymphocytes in either sector. Hippocampal injury grade as determined by the quantification of cells with features of acute cell death in the CA3 sector (Fig. 1D–I) was 56.1 ± 12.9 

Intercellular communication via cytokine upregulation in the injured hippocampus of SAMR1 mice

Although IFN-γ was once believed to be produced exclusively by T cells and NK cells, cells of the monocyte/macrophage lineage turned out to produce IFN-γ in the early phase of host response to infectious agents (Gessani and Belardelli, 1998). There is emerging evidence that microglia has capacity to produce IFN-γ. It has been reported that primary cultured microglia produce IFN-γ mRNA at 72 h following stimulation with interleukin (IL)-12 and/or IL-18 (Kawanokuchi et al., 2006). Microglia

Conflict of interest

The authors declare that they have no conflicts of interest.

Acknowledgments

We thank Dr. Yoshihito Tokita (Aichi Human Service Center) for technical support. This study was supported by Grants-in-Aid for Scientific Research from the Japan Society for the Promotion of Science (Contract Grant Nos.: 20790319 and 22790392 to SHI and 18590396 and 21590458 to AS) and the Research Grant C from Kansai Medical University (to MI).

References (82)

  • S. Hasegawa et al.

    Enhanced cell-to-cell contacts between activated microglia and pyramidal cell dendrites following kainic acid-induced neurotoxicity in the hippocampus

    J. Neuroimmunol.

    (2007)
  • J. Hung et al.

    Activation of microglia by neuronal activity: results from a new in vitro paradigm based on neuronal-silicon interfacing technology

    Brain Behav. Immun.

    (2010)
  • M.B. Jorgensen et al.

    Microglial and astroglial reactions to ischemic and kainic acid-induced lesions of the adult rat hippocampus

    Exp. Neurol.

    (1993)
  • S.Y. Kim et al.

    Osteopontin in kainic acid-induced microglial reactions in the rat brain

    Mol. Cells

    (2002)
  • T.B. Kirkwood

    Understanding the odd science of aging

    Cell

    (2005)
  • N. Kumagai et al.

    Involvement of pro-inflammatory cytokines and microglia in an age-associated neurodegeneration model, the SAMP10 mouse

    Brain Res.

    (2007)
  • S. Moriguchi et al.

    Potentiation of NMDA receptor-mediated synaptic responses by microglia

    Brain Res. Mol. Brain Res.

    (2003)
  • J.V. Nadler et al.

    Loss and reacquisition of hippocampal synapses after selective destruction of CA3–CA4 afferents with kainic acid

    Brain Res.

    (1980)
  • C.N. Nagineni et al.

    IL-11 expression in retinal and corneal cells is regulated by interferon-gamma

    Biochem. Biophys. Res. Commun.

    (2010)
  • L. Ottonello et al.

    CCL3 (MIP-1alpha) induces in vitro migration of GM-CSF-primed human neutrophils via CCR5-dependent activation of ERK 1/2

    Cell. Signal.

    (2005)
  • R.M. Raices et al.

    A synergistic role for IL-1beta and TNFalpha in monocyte-derived IFNgamma inducing activity

    Cytokine

    (2008)
  • D. Seilhean et al.

    Basic pathology of the central nervous system

  • A. Shimada

    Age-dependent cerebral atrophy and cognitive dysfunction in SAMP10 mice

    Neurobiol. Aging

    (1999)
  • A. Shimada et al.

    Localization of atrophy-prone areas in the aging mouse brain: comparison between the brain atrophy model SAM-P/10 and the normal control SAM-R/1

    Neuroscience

    (1994)
  • A. Shimada et al.

    Age-related deterioration in conditional avoidance task in the SAM-P/10 mouse, an animal model of spontaneous brain atrophy

    Brain Res.

    (1993)
  • R.L. Sprott et al.

    Historical development of animal models of aging

  • W.J. Streit

    Microglia and macrophages in the developing CNS

    Neurotoxicology

    (2001)
  • T. Takeda et al.

    A new murine model of accelerated senescence

    Mech. Ageing Dev.

    (1981)
  • J.H. Ye et al.

    Kainate-activated currents in the ventral tegmental area of neonatal rats are modulated by interleukin-2

    Brain Res.

    (2005)
  • W. Zhao et al.

    Differential expression of intracellular and secreted osteopontin isoforms by murine macrophages in response to toll-like receptor agonists

    J. Biol. Chem.

    (2010)
  • S. Alboni et al.

    Interleukin 18 in the CNS

    J. Neuroinflammation

    (2010)
  • R. Alonso et al.

    Interleukin-2 modulates evoked release of [3H] dopamine in rat cultured mesencephalic cells

    J. Neurochem.

    (1993)
  • D.N. Amin et al.

    Expression and role of CXCL10 during the encephalitic stage of experimental and clinical African trypanosomiasis

    J. Infect. Dis.

    (2009)
  • R.N. Auer et al.

    Hypoxia and related conditions

  • A. Bessis et al.

    Microglial control of neuronal death and synaptic properties

    Glia

    (2007)
  • D.J. Carr et al.

    Effect of anti-CXCL10 monoclonal antibody on herpes simplex virus type 1 keratitis and retinal infection

    J. Virol.

    (2003)
  • K. Chen et al.

    Induction of the formyl peptide receptor 2 in microglia by IFN-gamma and synergy with CD40 ligand

    J. Immunol.

    (2007)
  • Y. Chiba et al.

    The senescence-accelerated mouse (SAM): a higher oxidative stress and age-dependent degenerative diseases model

    Neurochem. Res.

    (2009)
  • J.R. Conde et al.

    Microglia in the aging brain

    J. Neuropathol. Exp. Neurol.

    (2006)
  • D.L. Dickstein et al.

    Changes in the structural complexity of the aged brain

    Aging Cell

    (2007)
  • R.N. Dilger et al.

    Aging, microglial cell priming, and the discordant central inflammatory response to signals from the peripheral immune system

    J. Leukoc. Biol.

    (2008)
  • Cited by (21)

    • Delayed microglial activation associated with the resolution of neuroinflammation in a mouse model of sublethal endotoxemia-induced systemic inflammation

      2022, Toxicology Reports
      Citation Excerpt :

      We hypothesized that microglia are activated in response to astrocyte-induced changes in the brain microenvironment, which includes alterations in cytokine concentrations, and thus, exert pro-inflammatory or anti-inflammatory functions. Microglia are brain resident macrophages [16,17], and therefore are considered to be involved in a variety of immune responses exerted by the brain [18,19]. Peripheral macrophages alter their phenotype depending on the microenvironmental status: M1 polarization in which macrophages function as a pro-inflammatory driver versus M2 polarization in which macrophages function as an anti-inflammatory driver [20,21].

    • Psychoneuroimmunology goes East: Development of the PNIRS<inf>China</inf> affiliate and its expansion into PNIRS<inf>Asia-Pacific</inf>

      2020, Brain, Behavior, and Immunity
      Citation Excerpt :

      The SAMP10 mouse strain exhibits microglial dystrophy from a young age and lacks the ability to construct a neuroprotective cytokine-mediated glia-neuronal network. This causes the mice to be vulnerable to age-related cognitive impairments (Hasegawa-Ishii et al, 2011). Japanese pathologists also invented a novel bone marrow transplantation method called intra-bone marrow-bone marrow transplantation (IBM-BMT) (Kushida et al, 2001), which led to the determination of the intracranial histological architecture supporting brain-immune cell-cell interactions (Hasegawa-Ishii et al, 2013).

    • Anti-inflammatory effects of thymoquinone in activated BV-2 microglial cells

      2015, Journal of Neuroimmunology
      Citation Excerpt :

      In resting BV-2 cells, osteopontin was the highest protein expressed, which could be a function of immortality as it is consistently reported in aggressive malignant glioblastomas (Jang et al., 2006; Matusan-Ilijas et al., 2008), in cerebrospinal fluid samples from glioma cancer patients (Yamaguchi et al., 2013) and its concentration is associated with invasiveness potential (Jan et al., 2010). Although very little is known about a role of osteopontin in primary microglia cells, it has been suggested to be a contributing factor to protection of nigral dopaminergic neurons somehow involving integrin and CD44 receptor interactions that promote neuronal remodeling (Hasegawa-Ishii et al., 2011; Liao et al., 2011) and survival (Del Rio et al., 2011). In this study, its quantity did not vary with LPS activation indicating its function is likely not related to inflammation.

    • Selective localization of bone marrow-derived ramified cells in the brain adjacent to the attachments of choroid plexus

      2013, Brain, Behavior, and Immunity
      Citation Excerpt :

      Based on their Iba-1 expression profile, ramified marrow-derived cells in the brain could be either microglia or brain dendritic cells (Bulloch et al., 2008; Prodinger et al., 2011). Given that microglia, astrocytes and neurons can establish cell-to-cell interactions by utilizing a variety of cytokines as mediators (Hasegawa-Ishii et al., 2011b), ramified marrow-derived cells may potentially interact with neurons and glia via cytokines. As well as under non-inflammatory conditions, ramified marrow-derived cells might also play an important role under inflammatory conditions.

    • Lymphocytes in neuroprotection, cognition and emotion: Is intolerance really the answer?

      2011, Brain, Behavior, and Immunity
      Citation Excerpt :

      In vitro IFN-γ enables microglia to remove glutamate without evoking inflammatory mediators (Shaked et al., 2005), but in vivo release of the other mediators would be expected unless regulatory cytokines were also present. The dilemma may be resolved by the fact that IFN-γ is made locally by microglia, and there is no need to suppose a requirement of Th1 T cells to supply this mediator to the injured brain (Hasegawa-Ishii et al., 2010). Interestingly, paroxetine and sertraline may inhibit microglial activation through inhibition of IFN-γ-mediated effects (Horikawa et al., 2010).

    View all citing articles on Scopus
    View full text