Innate immunity and cellular senescence: The ... - Wiley Online Library

8 downloads 0 Views 330KB Size Report
Jan 12, 2018 - Sagiv A, Biran A, Yon M, Simon J, Lowe SW, Krizhanovsky V. ... Winner B, Kohl Z, Gage FH. .... Henry CJ, Huang Y, Wynne AM, Godbout JP.
DOI: 10.1002/JLB.3MR0118-003R

REVIEW

Innate immunity and cellular senescence: The good and the bad in the developmental and aged brain Antonietta Santoro1

Chiara Carmela Spinelli2

Stefania Lucia Nori1

Mario Capunzo1

Stefania Martucciello3

Annibale Alessandro Puca1,2 ∗

Elena Ciaglia1 ∗ 1 Department of Medicine, Surgery and Dentistry

“Scuola Medica Salernitana,” University of Salerno, Via Salvatore Allende, Baronissi, Italy 2 Cardiovascular Research Unit, IRCCS Multi-

Medica, Milan, Italy

Abstract Ongoing studies evidence cellular senescence in undifferentiated and specialized cells from tissues of all ages. Although it is believed that senescence plays a wider role in several stress responses in the mature age, its participation in certain physiological and pathological processes

3 Department of Chemistry and Biology, Univer-

sity of Salerno, Fisciano, Italy

throughout life is coming to light. The “senescence machinery” has been observed in all brain cell populations, including components of innate immunity (e.g., microglia and astrocytes). As the ben-

Correspondence Elena Ciaglia, Department of Medicine, Surgery and Dentistry “Scuola Medica Salernitana”, University of Salerno, Via Salvatore Allende, 84081 Baronissi, Salerno, Italy. Email: [email protected] ∗ Annibale Alessandro Puca and Elena Ciaglia are

eficial versus detrimental implications of senescence is an open question, we aimed to analyze the contribution of immune responses in regulatory mechanisms governing its distinct functions in healthy (development, organogenesis, danger patrolling events) and diseased brain (glioma, neuroinflammation, neurodeneration), and the putative connection between cellular and molecular events governing the 2 states. Particularly this review offers new insights into the complex roles of senescence both as a chronological event as age advances, and as a molecular mechanism of brain

considered co-last authors.

homeostasis through the important contribution of innate immune responses and their crosstalk with neighboring cells in brain parenchyma. We also highlight the impact of the recently described glymphatic system and brain lymphatic vasculature in the interplay between peripheral and central immune surveillance and its potential implication during aging. This will open new ways to understand brain development, its deterioration during aging, and the occurrence of several oncological and neurodegenerative diseases. KEYWORDS

brain aging, cellular senescence, development, immune cells, inflammation, neurodegeneration

1

INTRODUCTION

cause or the consequence of a given pathology, senescent cells have been found in a large number of diseases1 where they share features of

From its discovery, cellular senescence has always been linked with

irreversible cell cycle arrest and phenotypic alterations, including chro-

aging and age-related disorders. Without addressing if they were the

matin and secretome modifications. When cells become senescent, they undergo profound changes,

Abbreviations: AD, Alzheimer's disease; AMD, age-related macular degeneration; APP, amyloid precursor protein; ARF, ADP-ribosylation factor; BBB, blood-brain barrier; CDK, cyclin-dependent kinase; CDKN2A, cyclin-dependent kinase inhibitor 2A; CSF, cerebrospinal fluid; CSF1, colony-stimulating factor 1; DDR, DNA damage response; DG, dentate gyrus; GSC, glioma stem cell; HCC, hepatocellular carcinoma; IKK𝛽, I𝜅B kinase-𝛽; MM, multiple myeloma; MMP, matrix metalloproteinase; NEP, neuroepithelial progenitor cells; NPC, neuronal progenitor cell; NSC, neural stem cell; OPC, oligodendrocyte precursor cell; PD, Parkinson's disease; PTEN, phosphatase and tensin homolog; RAGE, receptor for advanced glycoxidation end-products; ROS, reactive oxygen species; SA-𝛽-Gal, senescence-associated-𝛽-galactosidase; SASP, senescence-associated secretory phenotype; STING, stimulator of interferon genes; SVZ, subventricular zone; VEGF, vascular endothelial growth factor

Received: 3 January 2018

Revised: 12 January 2018

J Leukoc Biol. 2018;103:509–524.

including the acquisition of a senescence-associated secretory phenotype (SASP), one of the elements distinguishing senescent from quiescent cells.2 Senescent cells secrete proteins, including cytokines and chemokines with proinflammatory properties (CCL2/MCP-1, TNF-𝛼, IFN-𝛾, IL-6, IL-8), growth factors (TGF-𝛽, HGF), and proteases (MMP1/3/10/13, where MMP is matrix metalloproteinase) that have autocrine and paracrine activities, mainly allowing crosstalk between senescent cells and neighboring cells. This local and systemic alteration of tissue milieus leads to the detection, elimination, and rapid

Accepted: 12 January 2018 www.jleukbio.org

c 2018 Society for Leukocyte Biology

509

510

SANTORO ET AL .

replacement of senescent cells by the homeostatic protective system

of age-related pathologies,1 little is known about the induction of

of each tissue, but also to tumor formation and some age-related dis-

cellular senescence and the SASP in the brain. Here, environmental

orders when not properly regulated.1

and inner stressors may act in part by eliciting senescence and the

Senescent cells are detectable in the mammalian brain, specifically

SASP machinery within nonneuronal glial cells, thus contributing

in replication competent glial cells, where they could contribute to neu-

to the characteristic decline in neuronal integrity that occurs dur-

rodegeneration by secreting pro-inflammatory SASP factors and/or

ing brain aging. In this review, starting from a brief description of

deregulating cellular crosstalk useful for the structural and functional

the emerging role of senescence in different biological processes

neuron-glial interaction that maintains neuronal homeostasis.3 Senes-

(development, wound repair, oncogenesis, etc.) and of the reciprocal

cence markers have been observed in all the different cell popula-

influence with innate immunity in peripheral tissues, we will discuss

tions of the brain, including neurons and components of innate immu-

new emerging evidence indicating the existence of such a mechanism

nity, and they correlate with normal or pathological aging conditions.4

in developmental processes and in age-associated pathologies of the

Indeed, immune responses in the CNS are common, despite its per-

brain, in which microglia is recognized to be an essential player in

ception as a site of immune privilege.5 These responses are mediated

maintaining homeostasis.

by resident microglia and astrocytes, which are innate immune cells engaged in significant crosstalk with CNS-infiltrating T cells and other components of the peripheral innate immune system. Microglia and astrocytes, by interacting simultaneously with the peripheral immune system and typical brain cells, can lead to the resolution of infection, neurodegeneration, or neural repair and tissue development depend-

2 PERIPHERAL FEATURES OF SENESCENCE AND ITS ROLE IN NORMAL AND PATHOLOGICAL CONDITIONS

ing on the context.5 Very recently, Ablasser and colleagues6 illustrated how the entire

The concept of “cellular senescence” (or just “senescence”) was

process of senescence and the fast-acting, but unspecialized cells

originally formulated to describe the irreversible growth arrest of

of the innate immune system (macrophages, neutrophils, NK cells,

human fibroblasts after a finite number of divisions in culture.11

mast cells, etc.), share an overlapping mechanism of recognition of

Later, it was discovered that “replicative senescence” is linked to

aberrant self-molecule and microbial products, respectively. Follow-

telomere shortening resulting from consecutive cycles of DNA

ing defects in the integrity of the nuclear envelope, senescent cells

replication and cell division.12 Subsequent studies identified a

sense their own DNA, which activates the cyclic GMP-AMP synthase

“stress-induced senescence” independent from telomere erosion,

(cGAS)/stimulator of IFN genes (STING) pathway, a member of the

which was activated in response to a wide variety of stimuli, such

pattern recognition receptors, to regulate and facilitate the secretion

as DNA damage, oxidative stress, and activation of oncogenes.13–15

of inflammatory mediators of the SASP in an autocrine as well as a

Oncogene-induced senescence is due to activation of oncogenes,

paracrine manner by immune cells in vivo. These help the clearance

such as K-RAS, B-RAF, phosphatase and tensin homolog (PTEN), and

of aberrant senescent cells in various contexts of senescence, such

NF1, and is characterized by the derepression of the cyclin-dependent

oxidative stress, oncogene signaling, and irradiation, rather than fight-

kinase inhibitor 2A (CDKN2A) locus. Activation of the CDKN2A locus

ing off the pathogens.6 The study shows that DNA sensing through

produces the tumor suppressor p16 and ADP-ribosylation factor

the cGAS-STING pathway is an important regulator of senescence,

(ARF), causing growth arrest. This locus is repressed in young tissue

and the release of inflammatory mediators could serve as a surveil-

by polycomb group-mediated H3K27 methylation and H2A-K119

lance system that protects the organism against neoplastic cells and

ubiquitination, but becomes derepressed with aging or after tissue

in those conditions in which a programmed macrophage-mediated

damage.16 Oncogene activation has been reported to modulate

clearance of senescent cells is established, such as in organogenesis7

different pathways, many of which activate p53 and converge in

or tissue repair and fibrosis resolution.8 However, on the other face

the activation of the cyclin-dependent kinase (CDK) inhibitors p16,

of the coin, ones evoked and not self-resolving, the inflammatory

p15, p21, and p27, resulting in growth arrest.1 Telomere attrition

response can also contribute to the loss of tissue structure and func-

is sensed by cells as DNA damage, and like external DNA-damaging

tion by creating a chronic pro-inflammatory milieu leading to neurode-

agents, triggers the activation of the DNA damage response (DDR).

generation and cancer.9 Therefore, understanding the inner crosstalk

The DDR actives the ataxia-telangiectasia mutated and ataxia telang-

between innate immunity and cellular senescence might serve to iden-

iectasia and Rad3-related pathways, which in turn active the kinases

tify new drug targets to tackle diseases characterized by altered recip-

CDK2 and CDK1, respectively, determining p53-dependent cell

rocal interplay between different cell populations in organs, especially

cycle arrest.17 Reactive oxygen species (ROS) produced by exoge-

the CNS.

nous (such as radiation and smoke) and endogenous (mitochondria,

The occurrence of age-related neuropathology is increasing

RAS-RAF-MEK-ERK pathway) sources also induce senescence via

exponentially and has become a major public health concern in

mitochondrial and nonmitochondrial pathways, converging on p53,

recent years.10 Neuronal damage is believed to involve the induc-

pRB, p16, and p21,18 considered the main molecular players regulating

tion of neuroinflammatory events as a consequence of glial cell

cell senescence.

activation. Although cell senescence and the related inflammatory

Senescent cells are detected during embryo tissue development in

state in peripheral tissues has been causally linked to a number

highly differentiated and less-specialized cell types as well as during

511

SANTORO ET AL .

tissue repair and tumor suppression responses.19 Aging accelerates

by also inducing T-cell senescence, probably in a paracrine manner

the rate of senescent cell accumulation, as evidenced by a progressive

by a bystander effect. As regarding the innate immune compartment,

increase in some senescence markers. The senescence phenotype is

following inflammatory stimuli the onset of cellular senescence also

indeed characterized by a collection of markers, which can be used for

influences the polarization of macrophages, which play a critical role

its identification in vivo and in vitro. Not all these markers are exclu-

in innate immunity, acting as sentinels to fight pathogens, promoting

sively specific for senescent cells, which in turn do not simultaneously

wound healing, and orchestrating the development of the specific

express all of the senescence markers. To summarize, cultured senes-

acquired immune response. Specifically, deficiency of p16INK4a and

cent cells change their morphology, becoming much larger, flat, and

p14/p19ARF senescence markers was responsible for the polar-

vacuolized. Further, they are nonproliferating cells and do not show

ization both of murine bone marrow-derived macrophages27 and

BrdU incorporation or Ki67 expression. The most used method is his-

adipose tissue macrophages28 toward an alternatively activated anti-

tochemical detection of senescence-associated-𝛽-galactosidase (SA-𝛽-

inflammatory state, as suggested by the up-regulation of genes asso-

Gal), which reflects increased lysosomal content. Other conventional

ciated with the M2 phenotype, such as ARG1 (arginase-1), and Ym1/2

markers are p16, ARF, p53, p21, p15, p27, and hypophosphorylated

compared with wild-type cells. Accordingly, in the in vivo setting, aging

retinoblastoma (RB). 𝛾-H2AX (a phosphorylated form of the histone

affects many aspects of the cellular function of macrophages by mainly

variant H2AX) is a most sensitive marker of dsDNA breaks and telom-

impairing phagocytosis, removal of debris at the site of injury, wound

ere shortening, and the number of 𝛾-H2AX-positive foci increases in

repair, and tissue regeneration in both humans and rodents. Aging may

damaged and senescent cells.20 DDR is associated with the appear-

impair phagocytosis of myelin debris by macrophages, a key prerequi-

ance of senescence-associated heterochromatin foci , which contain

site step for efficient myelin regeneration (remyelination), and accord-

di- or tri-methylated lysine 9 of histone H3 (H3K9Me2/3), a histone

ingly Ruckh et al. elegantly demonstrated that monocytes from young

proteins.21

mice have the potential to boost remyelination in old mice.29 In con-

H2A variant (macroH2A), and heterochromatin protein 1

p38MAPK

and

trast, in a mouse model of age-related macular degeneration (AMD),

NF-𝜅B) are responsible for the expression of a number of senescence-

macrophages from old mice were unable to inhibit angiogenesis fol-

associated transcripts, while hyporeplicative senescent cells are

lowing laser injury to the retina when in the presence of IL-10.30 The

metabolically active.22

loss of antiangiogenic function of senescent macrophages was likely

Activation of damage-sensing signaling pathways (e.g.,

Cells with features of senescence have been identified during mam-

caused by down-regulation of FasL, suggesting that old macrophages

malian embryonic development at multiple locations as a phenomenon

might actually promote abnormal angiogenesis, as seen in diseases of

that contributes to remodeling and organogenesis. In this case, cells

aging such as AMD, the leading cause of blindness in people over 50

are recognized and eliminated by macrophages to efficiently remodel

years of age, and certain cancers. Finally, known for their cytotoxic

and sculpt tissues and organs.23 A classic example of senescence-

and immunoregulatory properties,31 NK cells are prone to undergo a

ducts.23

Fur-

senescence program. To date, this has been described only in a well-

ther recent studies have demonstrated the influence of senescence

established physiological setting. Specifically, in early pregnancy, the

driven sculpting is the regression of female Wolffian

repair.25

induction of NK senescence by DDR signaling generates the sustained

Senescent cells have been identified at different sites in the mouse

SASP secretory phenotype that can promote vascular remodeling

embryo, including the apical ectoderma ridge, the neuronal roof plate,

effects.23 Of note, these innate immune cells have been shown to elim-

the mesonephros, and the endolymphatic sac. These cells are nonpro-

inate cells carrying allogeneic mitochondrial DNA.32 Nevertheless, it

liferative, SA-𝛽-Gal-positive cells expressing p21, p15, and mediators

remains unclear whether immunosurveillance might efficiently antag-

of SASP, but they do not show DNA damage and do not depend on p53

onize the accumulation of mitochondrial DNA mutations normally

and p16 expression.7,23

associated with—and generally viewed as an etiological determinant

in embryonic

development,7,23

wound

healing,24

and tissue

of—aging.33

2.1

Cellular senescence and immune cell functions

Through the important contribution of immune cells, cellular senescence can control tissue repair and the maintenance of organ integrity.

Features of senescence, particularly the telomere-dependent ones,

The exponential accumulation of senescent cells with age can result

are widespread also throughout the immune system where intrinsic

from an increase in senescent cell production and/or decrease in

senescence-inducing actors could drive or tune the terminal differenti-

cell clearance, which usually occurs under physiological states. Effi-

ation program of several immune cell types, such as megakaryocytes, T

cient clearance of premalignant senescent hepatic cells through an

cells, macrophages, and NK cells. Senescence ensures proper terminal

immune response has been documented by Lowe and colleagues.34

differentiation of megakaryocytic cells, myeloid-derived immune cells

This depends on an intact CD4(+) T-cell-mediated adaptive immune

from which blood platelets originate, supporting normal and pathologi-

response, which, however, requires monocytes/macrophages to fully

cal wound healing and immune responses.26

Further, markers of senes-

execute the clearance of senescent hepatocytes. Impaired immune

cence, including CDKN2A, are also involved in the decision by mono-

surveillance of premalignant senescent hepatocytes results in the

cytes to differentiate toward inflammatory M1 macrophages, which in

development of murine hepatocellular carcinomas (HCCs), showing

turn will participate in T-lymphocyte polarization toward senescence-

that senescence surveillance is important for tumor suppression in

inducing Th1 cells. Notably CD4+CD25HiFoxP3+ Treg cells exercise

vivo. Indeed, in response to acute or chronic tissue injury, innate and

their immunosuppressive functions and limit inflammatory responses

adaptive immune cells greatly contribute to the maintenance of tissue

512

SANTORO ET AL .

homeostasis and help ensure the long lifespan of multicellular organisms. Injured cells within altered tissues/organs may become function-

3 SENESCENCE IN THE BRAIN AND ITS PHYSIOLOGICAL ROLE

ally deficient, leading to either cell death or the onset of cellular senescence. In the latter case, induction of senescence not only prevents the

3.1

Senescence onset in different brain cell types

potential proliferation and transformation of damaged/altered cells, but also favors tissue repair through the production of SASP factors1

The presence of senescence in the neural roof plate and in the neural

that function as chemoattractants mainly for NK cells (such as IL-

tube of mouse embryo reveals the importance that such a programmed

15 and CCL2) and macrophages (such as CSF1 and CCL2), innate

mechanism has in the development of the CNS.7 The mammalian neu-

immune cells endowed with a well-known cell clearance activity. These

ral tube is a monolayer of pseudostratified neuroepithelial progenitor

innate immune cells mediate an external quality-control mechanism

cells (NEPs), which move their nuclei up and down the apical–basal

for stressed cells, commonly known as immunosurveillance. Regard-

axis during the cell cycle (interkinetic nuclear migration). Initially, NEPs

less of the mechanism, senescent cells usually up-regulated the NK cell

undergo symmetric division to generate and expand the neural stem

activating receptor NK group 2 member D (NKG2D) and DNAM1 lig-

cell (NSCs) pool. After their early expansion—approximately at mouse

ands, which belong to a family of stress-inducible ligands, an impor-

embryonic day 11—NSCs begin to divide asymmetrically, forming a

tant component of the front line immune defense against infectious

daughter and a differentiating cell. In the later stage of this neurogenic

diseases and malignancies. Upon receptor activation, NK cells can then

phase, intermediate neuronal progenitor cells are formed, migrate to

specifically induce the death of senescent cells through their cytolytic

the subventricular zone (SVZ), and divide symmetrically to produce

machinery. Remarkably, a role for NK cells in the immune surveillance

neurons. Finally, gliogenesis begins at embryonic day 17, generating

of senescent cells has been pointed out in liver fibrosis,35 HCC,36 multi-

astrocytes and oligodendrocytes.46 Markers of senescence have been

ple myeloma (MM),37 and glioma cells stressed by dysregulation of the

identified in vivo in the neural tube up to this stage of the neurogenic

mevalonate pathway.38

phase.7 NSCs persist in the subgranular zone of the dentate gyrus

With a similar mechanism, during liver fibrosis, p53-expressing

(DG) of the hippocampus and in the SVZ of the lateral ventricles in the

senescent liver satellite cells skewed the polarization of resident Kupf-

mammalian adult brain.47 The NSCs ensure brain homeostasis, repair,

fer macrophages and freshly infiltrated macrophages toward the pro-

and neuroregeneration in the adult, but undergo proliferation and dif-

which displays senolytic activity. In

ferentiation decline with age. The decreased neurogenesis correlates

the same way, F4/80+ macrophages are key players in the clear-

and contributes to age-related disorders in the human and in mouse

ance of mouse uterine senescent cells in order to maintain postpar-

model,48,49 and it is thought to be due to the senescence of progenitor

tum uterine function.40 Cancer is another common disease of aging,

cells. Diverse stimuli can induce senescence in NSCs in vitro, includ-

and it is now clear that dysregulations in key mediators of senes-

ing DNA damage,50,51 oxidative stress,52 and activated oncogenes.53

cence (e.g., telomerase, p16INK4 , p53, RB) may influence human cancer

Several in vivo studies have supported this hypothesis, demonstrat-

risk. Paradoxically, if aging leads to decreased regenerative capacity in

ing age-dependent increase in telomere shortening, p16INK4a expres-

the brain, conversely it increases the risk of tumorigenesis.41 While

sion, and ROS accumulation in mouse NSCs.54–56 NSCs are therefore

considered to be an antitumor mechanism for cancer prevention,42

the physiological reservoir for neurons, astrocytes, and oligondendro-

senescent cells, through the release of different cytokines, includ-

cytes, whereas microglial cells are the only cells in the brain deriving

ing IL-1 and IL-6, may also influence other cells in an autocrine

from hematopoietic precursors of the yolk sac; once migrated, they are

and paracrine fashion, both enhancing and dampening antitumor

self-renewed and do not require maintenance upon the recruitment of

responses. Indeed, the ability of the SASP to induce an immuno-

bone marrow-derived cells.57

inflammatory M1

phenotype,39

suppressive immune cell environment has recently been shown in a

Neurons are terminally differentiated, postmitotic cells that

mouse model of PTEN loss-induced cellular senescence in prostates

remain functional for the lifetime of the organism. Although these

with prostatic intraepithelial neoplasia.43 Senescence-recruited imma-

cells do not undergo replicative senescence, recent studies revealed

ture myeloid cells and macrophages cleared precancerous senescent

that neurons can show a stress-induced senescence-like phenotype.

cells, but later they were responsible for senescence-induced tumor

Increased 𝛽-galactosidase activity and mitochondrial dysfunction

activity.44

were observed in hippocampal neurons in long-term cultures in

Further, MM causes clonal T-cell immunosenescence and, accordingly,

vitro,58 and 𝛽-galactosidase-positive staining was detected also in vivo

the cells exhibited a senescent secretory effector phenotype: KLRG-

in the hippocampus of aging rat.59 Furthermore, in response to DNA

1+/CD57+/CD160+/CD28–. Tumor-induced senescent TCD8+ cells

damage, purkinje, hippocampal, and cortical mouse neurons exhibit

with suppressor function have been defined a potential form of tumor

different markers of senescence, including heterochromatinization,

immune evasion.45 This is in contrast with the above-discussed find-

synthesis of proinflammatory interleukins, and high b-galactosidase

ing of efficient immune cell clearance of senescent cells as an extrin-

activity. This p21-dependent phenotype is aggravated in the aged

sic antitumor barrier, suggesting a possible dual role of cellular senes-

mouse and attenuated by caloric restriction.60 Thus, senescent neu-

cence in tumor prevention and tumor progression. Future studies are

rons might contribute to brain aging as an important source of chronic

thus needed to fully understand which molecular and cellular factors

pro-inflammatory and pro-oxidant signaling.

growth promotion in mice, leading to inhibition of NK cell

secreted or overexpressed by senescent cells may determine the final outcome of immune responses.

Neuroglial cells, including astrocytes, oligodendrocytes, and microglia, represent a large component of the mammalian brain.

513

SANTORO ET AL .

Astrocytes and oligodendrocytes have an important supporting role in

The most important and beneficial cell types for maintaining normal

neural functions. In particular, oligodendrocytes produce the myelin

brain function are microglial cells, large macrophages that account for

that surrounds the axon of nerve cells and facilitates the conduction

∼15% of the brain's cellularity. Microglia preserve brain homeostasis

of signals. Changes in oligodendrocyte morphology, including the

by rapidly responding for clearance of invading pathogens, tissue

swelling of cell processes and the presence of inclusion bodies, were

debris, synaptic stripping, and remodeling.9 However, as we age,

observed in human aged brain, and were correlated with the progres-

microglia undergo a process of immunosenescence, characterized by

sive demyelination of nerves during aging and neurodegenerative

changes in morphology and a more reactive/activated phenotype.79

diseases. Magnetic resonance imaging has confirmed the reduction of

In aged brain, microglia from human cerebral cortex exhibit deram-

white matter volume and structural integrity both during aging and

ification, spheroid formation, and fragmentation of processes.80

in neurodegenerative diseases.61 The progressive loss in remyelina-

Telomere shortening has been seen to occur in rat microglia, both in

tion efficiency occurs because of an impairment of oligodendrocyte

culture after repeated cell division and in an advancing state in vivo.81

precursor cell (OPC) recruitment and the subsequent differentia-

Morphological abnormalities associated with the switch from an

tion into remyelinating oligodendrocytes.62 Furthermore, rodent

amoeboid to a ramified morphology were also observed in microglial

OPCs escape from replicative senescence in vitro,63 but acquire an

cells isolated from neonatal mice and cultured from day 2 to 16 in

ECRG4-mediated senescence phenotype when grown with serum.64

vitro.82 Biochemical alterations were also found, such as a reduction

Astrocytes, the most present glial cells, constituting about 50%

of MMP-9 and glutamate release, an increase in NF𝜅B activation,

of CNS cells,65,66 are active players in brain homeostasis and central

and a decreased expression of TLR-2 and TLR-4, resembling the

innate immunity. They react to different insults and injuries through

activated as well as senescent phenotype.82 Other authors reported

a spectrum of molecular, cellular, and functional changes, defined as

morphological changes in terms of size, heterogeneous cytoplasmic

astrogliosis.67,68 Senescence hallmarks have been detected in cultured

content, IL-1 expression, and transcription in the mesial temporal

astrocytes in response to oxidative stress. They involve morphologi-

lobe of normal aged individuals,83 suggesting that altered cytokine

cal changes, expression of p16, p21, p53, and p53-binding protein 1

and TLR expression profiles are associated—but not necessarily in a

(53BP1), cell cycle arrest, reduction in telomere length, and secretion

cause-effect manner—with the morphological changes in microglia.

In vivo, it is conceivable

Sierra et al.84 demonstrated that in aging mouse microglia in vivo,

that their age-associated changes are more complex than previously

there was a characteristic presence of lipofuscin granules, decreased

described and involve either the expression of the same senescence

processes complexity, altered cytoplasmic granularity, and increased

markers of other kind of cells or the production of specific factors

mRNA expression of either pro-inflammatory (TNF-𝛼, IL-1𝛽, IL-6)

required for the crosstalk with the neighboring cells. Aged hypothala-

or anti-inflammatory (IL-10, TGF-𝛽1) cytokines.84 Furthermore,

mic astrocytes from brain-specific I𝜅B kinase-𝛽 (IKK𝛽) knockout mice

in healthy adult and aged mice, RNA sequencing followed by RT-

show alterations in intracellular signaling pathways involved in neu-

PCR and fluorescence dual in situ hybridization has revealed that

roinflammation, such as activation of NF𝜅B and its regulatory proteins

microglia have a distinct cluster of transcripts encoding proteins for

IKK𝛽,70

also influencing microglia-neuron crosstalk. Mechanistic stud-

sensing endogenous ligands and pathogens, named the sensome.85

ies in this genetic model further showed that IKK𝛽 and NF𝜅B activation

With aging, sensome transcripts appeared down-regulated, whereas

inhibited gonadotropin-releasing hormone, promoting aging.70 More-

transcripts involved in microbe recognition and host defense were

over, it has been proposed that the functional decline of aged brain may

up-regulated together with an increase in the expression of genes

be due, at least in part, to decreased trophic support, changes in synap-

that promote neuroprotection.85 These parallel increases in both

of pro-inflammatory cytokines (SASP

like).69

astrocytes.71,72

In

pro-inflammatory and anti-inflammatory factors suggests that, even

agreement with this hypothesis, more recent in vitro studies demon-

though in aged microglia basal levels of these proteins are higher than

strated that hypothalamic astrocytes from aged Wistar rats are able

young microglia, cells are still able to self-regulate switching from the

to remodel and change their neurochemical properties compared to

classical M1 state to M2, promoting resolution of inflammation and

newborn and adult rats.73

Age-related variations were observed in the

restoring homeostasis.86 From this point of view, it has been proposed

regulation of glutamatergic homeostasis (with a significant reduction

that with age, microglia could be in a chronic state of neuroinflam-

of glutamate production in aged astrocytes), glutathione biosynthe-

mation. This implies that cells are already “primed” and, as in the case

sis, amino acid profile, and glucose metabolism. Additionally, increased

of peripheral macrophages, a secondary triggering stimulus induces

expression and activation of NF𝜅B and PI3K/Akt was observed, cor-

a more rapidly and greeter response in terms of pro-inflammatory

roborating that during aging physiological changes of astrocytes could

cytokines. Accordingly, some in vivo studies demonstrated that sys-

modify their impact on microglia and neuron functions. Moreover,

temic treatment with LPS to aged mice elevates the level of cytokine

studies on rat and human postmortem brain samples suggest a direct

production compared to young animals.87,88 Frank et al.89 examined

communication of astrocytes with microglia in aged brain, since it

age-related mRNA alterations of cell surface MHC class II in the

has been found that during aging, astrocytes from hippocampal sub-

hippocampus of older (24-month old) and younger (3-month old) male

regions (CA1, CA2, and CA3 sectors) have increased expression of

F344xBN F1 rats. They found that older animals exhibited increased

glial fibrillary acidic protein and vimentin74–76 and show a hyper-

mRNA levels of MHC class II, CD86, CIITA, and IFN-𝛾, while exhibiting

tic efficacy, and an increase in ROS production in

trophic

morphology77

microglial functions.78

typical of an activated state that can modulate

a reduction of molecules down-regulating macrophage activation, such as IL-10 and CD200. These findings corroborate the hypothesis

514

SANTORO ET AL .

that normal brain aging is characterized by a shift toward a chronic, but

leptomeninges, and periventricular regions [reviewed in Ref. 107]. In

physiological, low-grade pro-inflammatory state, which we can define

the light of these results, in our opinion we should revisit our traditional

as “brain inflammaging”. Indeed, inflammaging is referred to as a sub-

concept of the brain as an immune-privileged place: rather, it is a terri-

clinical chronic inflammatory process90,91 characterized by peripheral

tory with a double and cooperative immunity. From the inside, resident

immune innate activation and significant changes in monocyte and

microglial cells and astrocytes would control inflammation and coordi-

macrophage functions as well as by a parallel decrease in peripheral

nate brain response to peripheral attacks through its crosstalk with T

naive T cells and an increase in late-stage differentiated memory T cells

cells. From the outside, naïve myeloid precursors would be necessary

with a reduced antigen receptor diversity.92,93 However, changes in the

in the context of brain inflammation, injury or infection in which, as in

basal pro-inflammatory state of innate immunity both in the periphery

other body tissues, the general immune system must be recruited.104

and CNS have some similarities, and peripheral immunosenescense

The existence of this spatial, temporal, and functional cooperation

may contribute to neuroinflammation by modulating glial cells toward

has been reinforced by the recent discovery and description of both

a more active pro-inflammatory state.94 As an environmental sensor,

the brain glial-dependent lymphatic system, named the “glymphatic

microglia interact with the surrounding cells, astrocytes, oligoden-

system”108,109 and meningeal lymphatic vessels.110 The former is

drocytes, and neurons, so changes in microglia during brain aging are

composed from a wide network of paravascular pathways and astro-

expected to influence the functions of the other CNS cells.

cytes that connect cerebrospinal fluid (CSF) and its circulation with

Emerging evidence highlights that inflammaging modifies neuron–

interstitial fluids and the brain parenchyma.108–110 Although the glym-

astrocyte–microglia crosstalk and that this interaction could be

phatic system has been considered the essential fluid drainage system

responsible for brain aging and, when impaired, for neurodegenera-

responsible for transferring interstitial fluids and solutes including

tive disease.95,96 Microglia as a key immune regulator of CNS involves

misfolded proteins such as A𝛽 from the brain parenchyma to the

also the interplay with neurons. In aged brain, it appears that some lig-

CSF,108,111 how CSF can leave the CNS and entry into the deep cervi-

ands of the neurons for microglia-specific receptors, such as CX3CR1

cal lymph nodes has been a subject of interest for decades. This has led

and CD200 protein expression, decreased with increasing age, con-

to the discovery of a lymphatic vessel network lining the dural sinuses

tributing to establishing a pro-inflammatory phenotype.96–98 Finally,

in the mouse brain meninges.112 These structures express all of the

microglial activation results in the release of soluble factors that

molecular hallmarks of lymphatic endothelial cells, are able to trans-

can prevent TNF-𝛼-induced apoptosis of mature oligodendrocytes in

port fluid and immune cells from the CSF, and are connected to the

vitro.99 However, we have to take into account that most of the exper-

deep cervical lymph nodes, from which antigens may potentially induce

iments were carried out in animal models of neurodegeneration and

an immune response.110 Meningeal lymphatics are now considered

in experimental conditions of extreme activation of microglia, so the

a novel path for immune cells to egress the CNS under physiological

significance of such alterations in physiological conditions during aging

and pathological conditions, and its description challenges the dogma

remains to be elucidated.

that the immune privilege of CNS is in part due to the absence of CNS

On the other hand, microglia play also a role in neural progenitor

lymphatic drainage. Indeed, it is plausible that meningeal lymphatic

cell differentiation because they release soluble factors that exhibit

endothelial cells may have unique properties that induce and control

distinct effects on the different aspects of neurogenesis upon the acti-

CNS tolerance by shaping immune responses, as recently described

vation of TLRs100 ; this has led recently to the concept of “senescence

for lymphatic endothelial cells,113,114 using a transgenic mouse model

spreading” and/or paracrine senescence,101,102 which opens new ques-

expressing vascular endothelial growth factor (VEGF)-C/D trap and

tions into how dysregulated microglia can influence NSCs senescence

displaying complete aplasia of the dural lymphatic vessels. Aspelund

in the brain.

and co-workers112 demonstrated that the dura mater's lymphatic ves-

Even though glia (predominantly microglia and astrocytes) rep-

sels are very sensitive to the inhibition of VEGF-C/D signaling, since in

resent the main components of immunesurveillance in the CNS, and

these mice macromolecule clearance from the brain was attenuated,

cells of adaptive immunity cannot pass unrestrictedly the blood-brain

and transport from the subarachnoid space into deep cervical lymph

barrier (BBB), adaptive immune cells recruited from blood have been

nodes was abolished. Then malfunction of these meningeal lymphatic

found in several CNS-related autoimmune pathologies, such as multi-

vessels could be a cause of different pathological brain conditions in

ple sclerosis.103 Potential gates for leukocyte entrance are beginning

which altered immunity plays a key role, such as multiple sclerosis,

to be found [reviewed in Ref. 104], but how immune cells can reach the

Alzheimer's disease (AD), and possibly other aging-related disorders.

brain parenchyma, if they can recirculate, and the spatial and temporal

A significant decline in CSF outflow via meningeal lymphatics was

requirements that allow them to pass without compromising BBB

observed in aged mice, implying that meningeal lymphatic network

integrity are still matters of debate. It has been proposed that recir-

could be targeted for age-associated neurological conditions.115,116

culation of infiltrating immune cells from the brain parenchyma might not occur under physiological conditions,105 and the blood-meningeal barrier—which does not comprise astrocytes—could be more permis-

3.2

Developmental senescence in the brain

sive than BBB to immune cells entrance, allowing them to circulate

The role of senescence mechanisms in the acquisition of a properly

within the meninges.106 Moreover, MHC class II- and B7-positive

function of innate immune cells in CNS and in neuronal circuits they

perivascular macrophages, which are distinct from microglial resident

established is coming to the light. Cytokines and their receptors are

APC cells, have been described in the choroid plexus, dura mater,

expressed physiologically in CNS cells and are important for many

515

SANTORO ET AL .

neurodevelopmental processes and for the regulation of several spe-

during synaptic refinement.133 An additional function of microglia

cific area of the brain.117,118 A large number of pro-inflammatory

includes a significant contribution to synaptic stripping or remodeling

cytokines (e.g., IL-1𝛽, TNF-𝛼, TGF-𝛽) are expressed at high levels in

events.125 Activated microglia can up-regulate MHC class II surface

the developing CNS but at very low (constitutive) levels in the adult

expression and release cytokines, chemokines, nitric oxide, and sev-

brain. Coinciding with the appearance of amoeboid microglia during

eral neurotrophins, which regulate developmental cell death. Finally,

early brain development, researchers have reported a naturally occur-

microglia are observed within the myelin tracts during early develop-

ring increase in these cytokines suggesting their physiological role in

ment and changes in white matter microglia have been associated with

the development of specific brain area where they have been shown

deficits in oligodendroglia progenitor cells or myelination.134

to maintain neuronal identity and glial differentiation, proliferation,

Beside microglia, also astrocytes and oligodendrocytes have been

and synaptic maturation.119 For example, it has been demonstrated

characterized for their role in neuronal development. In mice, astro-

that active inhibition of bone morphogenetic proteins signaling, which

genesis starts around embryonic age 18 (E18) and lasts until approx-

are members of the TGF-𝛽 cytokine superfamily, is required for nor-

imately postnatal day 7. Astrocytes have precursors of neuroepithe-

mal neural development in mice.120 In rodents, IL-1𝛽 is produced at

lial origin, the above-mentioned neuronal progenitor cell (NPC),135

detectable levels within the cortex from approximately E14 to P7

that transform into radial glia. At the end of neurogenesis, radial glia

where it contributed to astrogliosis and neovascularization and has a

cells can differentiate into mature astrocytes characterized by changes

peak in the late development of cerebellum that occurs from P2 to

in morphology, connectivity, and electrophysiological properties.136

P14121

and also within the hippocampus; IL-1𝛽 levels are increased

Mature astrocytes were present throughout much of the normal CNS

nearly 6-fold at birth when compared to adult hippocampus.122 IL-6

at 15 weeks of gestation, but they vary in density in different parts.137

is important for numerous developmental processes including vascu-

Although astrocytes appear much later in the brain during develop-

logenesis of microvessel endothelial cells in murine prenatal brain.123

ment, it has been described that perinatal inhibition of astrogenesis

IL-6 increases markedly in striatum, hippocampus, and cortex through-

resulted in a drastic reduction in the density and branching of corti-

out development, suggesting a neurotrophic role for this cytokine

cal blood vessels suggesting to play a part in postnatal developmental

within these brain regions.124 All these data suggest that elevated lev-

angiogenesis.138 They were shown to be actively involved in the for-

els of particular cytokines may coincide with important processes of

mation and stabilization of blood vessel of the retinal vasculature139

neurodevelopment in a brain region-dependent manner. As this pro-

related to their properties to express the VEGF.140 Moreover, many

inflammatory secretory phenotype commonly belongs to senescent

studies suggest that astrocytes are involved in the formation and in the

cells characterized by a SASP phenotype, we might speculate that

maintenance of BBB properties141 by producing secreted molecules

senescence might also occur in and/or implicated in the fine tuning of

important for the regulation of interactions between BBB components

brain immune cells to maintain the brain integrity in early development

such as endothelial cells and pericytes.142,143

stage and in tissue repair processes.

Oligodendrocytes are highly specialized neural cells whose function

There are several reports indicating that microglial cells contribute

is to myelinate CNS axons. They are cells that form the fatty sheaths

to fetal brain development.125 In rat microglial cells are observed

that cover axons of neurons passing through the brain. It is becoming

in the developing brain at embryonic day 16 (E16) around subcorti-

more evident that glia-glia crosstalk plays several important roles in

cal regions.126 Microglial development occurs in parallel with neuro-

brain function during development [reviewed in Ref. 65]. In particular,

genesis in most brain structures.127 They arise from hematopoietic

astrocytes facilitate each step of myelination, including OPC prolifera-

stem cells in the yolk sac during early embryogenesis that populate

tion, differentiation, initial oligodendrocyte-axon contact, and myelina-

the CNS.128,129 The initial colonization of the brain by microglia cor-

tion. Moreover, other soluble factors secreted by astrocytes have been

responds to the vascularization of the CNS. During this time, cell-

implicated in enhancing myelination.65

cell communication continues between microglia and vascular sprouts, directly after they enter the neuroepithelium and throughout CNS development, suggesting a potential role for microglia in blood vessel formation.130,131 Therefore, in the human brain, microglia colonization coincides not only with vascularization but also with radial glia forma-

4 PATHOLOGICAL SIGNIFICANCE OF SENESCENCE IN THE BRAIN

tion, neuronal migration, and myelination.125 It has been speculated that microglia perform specialized functions

As innate immune cell functions decline with aging,144 process termed

critical to the development of the brain. For example, microglia appear

immunosenescence,90 it is tempting to speculate that a parallel decline

to influence events associated with neuronal proliferation and differ-

in cell clearance mechanism might occur as culprit in the deleteri-

entiation during development: microglia have the capacity to influ-

ous accumulation of senescent cells with aging. The immunological

ence the differentiation of adult and embryonic neural precursor cells

“silent” removal of apoptotic and increasing number of senescent cells

and it has been demonstrated that precursors depleted of microglia

might be compromised and may contribute to the pro-inflammatory

decreased precursor proliferation and astrogenesis, and these deficits

phenotype as discussed above. This might explain the most critical

could be rescued when microglia were added back to the cultures.132

change in the aging innate immune system, which is the increase in the

Additional function of microglia includes a significant contribution to

pro-inflammatory cytokines IL-1, IL-6, IL-18, and TNF-𝛼 suggestive of

synapse formation: they may remove complement-labeled synapses

a SASP phenotype145 and the consequent low-grade inflammation,

516

SANTORO ET AL .

contributing to immunosenescence and age-related disease, especially

recently demonstrated that glioma stem cells (GSCs) express and in

neurodegenerative disorders. Indeed when senescent cells were

turn induce the secretion by monocytic myeloid-derived suppressor

removed from aged mice artificially, the animals lived longer and were

cells, of many SASP factors, that can act as mediators of intercellu-

healthier.146 This was also confirmed by pharmacological approach

lar communication to promote systemic tumor immune escape. In par-

demonstrating that the “senolytic” clearance of senescent cells can

ticular, GSC-derived molecules are endowed with an immunomodula-

reduce age-related phenotypes.147 In the brain, the detection and the

tory activity that results in the suppression of peripheral T-cell immune

clearance of invading microorganisms and senescent cells as well as

responses. This is achieved by the engagement of cells of myeloid

surplus neurotransmitters, aged and glycated proteins, in order to

origins,156 which by losing their protective tumor surveillance function

maintain a healthy environment for neuronal and glial cells, is largely

clearly became cellular enemy in gliomagenesis.

confined to the innate immune system. Neuronal and glial cells express TLRs as well as complement receptors and virtually all complement components can be locally produced in the brain, often in response to

4.2

Senescence in brain neurodegenerative diseases

injury or developmental cues148 or to finely orchestrate wound repair

Neurodegeneration is characterized by synaptic loss and neuronal

as for microglial cells.149 In this context, it is tempting to speculate that

death resulting in mental impairments and functionality. It is commonly

immunosurveillance mechanisms may be in place to preserve neuronal

accepted that chronic inflammation contributes to neuronal degen-

homeostasis, and that the aging-related failure of such systems, for

eration. Neural damage in the CNS preferentially involves the acti-

the inner progressive decline in immune effector function, may allow

vation of the resident glial cells: microglia and astrocytes.157 How-

the development of neurodegenerative disorders and at least under

ever, activated innate immunity in the brain is able to elicit a strong

some circumstances, tumor occurrence. In the brain, innate immunity

adaptive immunity reaction through accumulation of misfolded pro-

responds differently to all kind of attacks, which can be divided

teins inducing neuroinflammation and long-term, low-grade sustained

essentially into 4 groups: acute injuries comprising traumas, ischemic

inflammatory factors that stimulate chronically both innate and adap-

stroke, and so on; neurodegenerative chronic disease (AD; Parkinson's

tive immunity.158 In this context, both AD and PD are age-related neu-

disease [PD], amyotrophic lateral sclerosis, and multiple sclerosis);

rodegenerative pathologies characterized by accumulation and aggre-

brain tumors (gliomas, glioblastomas, etc.); and infections (Escherichia

gation of misfolded proteins: the A𝛽 peptide in the first case and the

coli, HIV, etc.).9 Here, we would provide an overview on the knowl-

𝛼-synuclein in the latter.

edge of the relationship between innate immunity, senescence, and

AD is a neurodegenerative disorder characterized by progressive

brain pathologies, focusing on brain aging-related neurodegenerative

loss of neuronal functions leading to cognitive and functional impair-

disease and cancer.

ment and memory loss. The main pathological hallmarks of AD are the abnormal cerebral production and accumulation of A𝛽 peptide—

4.1

generated by the proteolytic cleavage of the amyloid precursor pro-

Senescence and brain cancer

tein (APP) in neurons—and the deposition of neurofibrillary tan-

Age is a strong predictive factor in the occurrence of glioma, the

gles (composed of hyperphosphorylated tau protein) within the brain

most aggressive adult brain cancer.38,150 Little is known about how

parenchyma and blood vessels.

aging increases glioma malignancy151 but recent studies highlighted

In AD, A𝛽 is present in the extracellular environment in various

the contribution of a pool of dysregulated NPCs with an increased

aggregation states (A𝛽 oligomers, fibrils, and plaques).159 As A𝛽

tendency to undergo senescence as opposed to proliferation and self-

oligomers accumulate, they prime microglial cells promoting their acti-

renewal in response to growth signals. Indeed, genome-wide associ-

vation and thus generating a pro-inflammatory status that perpetuates

ation studies and candidate analysis have identified noncoding reg-

and completely changes extracellular environment. It has been demon-

ulatory polymorphisms near the CDKN2A locus, producing p16INK4

strated that A𝛽 oligomers can interact with several receptors including

and ARF, that affect the lifetime risk of glioma.152 Telomere main-

TLR2, TLR4, CD14, CD36, and CD47 both in vivo and in vitro.160–163

tenance has emerged as another important molecular feature with

More recent results have demonstrated that ablation of TNF-RI/RII

impacts on adult glioma susceptibility and prognosis,152,153 suggest-

expression in a mouse model of AD can enhance pathology,164

ing that telomere biology and induction of replicative senescence may

suggesting that increased expression of TLR receptors and inflamma-

have a role in gliomagenesis. In glioma, the loss of immunesurveillance,

tory mediators can contribute to disease progression. Moreover, some

due to “immunosenescence”, may contribute to age-related increases

studies in vitro revealed that microglia are able to clear fibrillar A𝛽 acti-

in glioma incidence. One recent study showed that the decreased pro-

vating phagocytosis via CD14, TLR4, TLR2, and 𝛼6𝛽1 integrin.165,166

duction of CD8+ T cells is associated with increased glioma malignancy

Microglial cells also produce intracellular and extracellular proteases

in both aged human patients and a knockout mouse model.154 Further

such as neprilysin- and insulin-degrading enzyme eliminating A𝛽

gliomas activate microglia, but inhibit their phagocytotic activity and

soluble oligomers.167 These observations suggest that microglia acti-

enhance expression of pro-migratory

metalloproteases.155

Undoubt-

vation could participate to limit disease progression; however, when

edly, the major contributing factor to glioma development and progres-

A𝛽 production exceeds microglial ability to clear the toxic oligomers,

sion is its ability to evade the immune system. To date, only 1 evidence

microglia enter in an unresolving circuit of inflammation that has

would suggest the occurrence of an aberrant senescent program as a

been defined “frustrated inflammation.”9 Results obtained in animal

al.156

model and in vitro cultures of murine microglia are corroborated by

strategy of immune evasion in brain cancer. Indeed, Domenis et

517

SANTORO ET AL .

studies in humans. Human primary microglial cells are activated with

bradykinesia, rigidity, postural instability, and gait imbalance. The

the treatment of A𝛽 oligomers and exhibit up-regulated mRNA and

motor symptoms are generally considered the consequence of the

protein expression of pro-inflammatory cytokines including IL-1𝛽,

loss of dopaminergic neurons in the substantia nigra pars compacta of

IL-6, MCP-1, TNF-𝛼, and the chemokines CXCR2, CCR3, CCR5, and

the midbrain. The main pathological hallmark of PD is the deposition

TGF-𝛽.168 The overexpression of such a kind of cytokine milieu was

of an intracellular fibrillar and misfolded protein named 𝛼-synuclein

also found in human cortex of postmortem AD brains and in the CSF of

(Lewy bodies) causing a complex immunopathogenic response leading

Noteworthy, microglia, astrocytes, endothelial cells,

to PD.181 The etiology of the disease is still unknown; several studies

and neurons also express the receptor for advanced glycoxidation

in disease animal models suggest that PD could be a T-cell-dependent

end-products (RAGE), a cell surface receptor that is able to interact

autoimmunity associated with neuronal death. It has been demon-

with A𝛽 peptide and oligomers.169 Blocking the interaction of A𝛽 with

strated that CD4+ T cell expressing altered levels of the dopamine

RAGE impaired the activation of microglia and reduced the production

receptor D3 favors acquisition of the Th1 inflammatory phenotype,182

of proinflammatory mediators [reviewed in Ref. 170]. RAGE is also

whereas other authors proposed that T cells specific for the neurome-

suggested to play a key role in the clearance of A𝛽 and to be involved in

lanin can activate neuromelanin-specific B cells leading to the produc-

apoE-mediated cellular signaling, a family of proteins widely expressed

tion of autoantibody. In any case, stimulated T cells are able to infiltrate

in brain cells that by binding their receptor activate different cell kind

into the brain and activate resting microglia and astrocytes,181,183,184

dependent signaling pathways regulating the whole cellular crosstalk

finally causing neuroinflammation. On the other hand, microglia are

in the brain.171,172

able to recognize and phagocytize extracellular 𝛼-synuclein aggregates

AD

patients.9,169

As astrocytes display pleiotropic functions beyond their involve-

and are also activated by neuronal cell death itself. It has been shown

ment in supporting cells of the brain, data are accumulating on the role

that released extracellular 𝛼-synuclein or protein aggregates deriv-

of astrocytes in AD pathogenesis and progression. AD patients show

ing from neuronal death can be internalized by microglia through a

hypertrophic astrocytes associated with changes in GABA signaling

lipid rafts-mediated mechanism in BV2 cells.169,185 In a primary mes-

and recycling, potassium buffering, and in cholinergic, purinergic, and

encephalic neuron-glia culture model of PD, 𝛼-synuclein is phagocy-

calcium signaling.173 This phenotype, referred to as “reactive astro-

tosed by microglia resulting in increased pro-inflammatory cytokine

It

and chemokine production and activation of NADPH oxidase,186 which

has been shown that in mouse models of AD (5xFAD and APP/PS1

is the main source of ROS production in activated microglia. Accord-

mice), reactive astrocytes in the DG dramatically up-regulate intracel-

ingly, enhanced levels of IL-1𝛽 and IL-6 were found in plasma, nigros-

lular GABA levels compared with astrocytes from age-matched wild-

triatal regions, and CSF of PD patients compared to healthy aged

cytes”, presents a morphology resembling senescence

GABA levels were inversely correlated with the dis-

subjects.187,188 The high pro-inflammatory activation of microglia can

tance to amyloid depositions, suggesting that both hypertrophy and

be ultimately responsible for the recruitment of CD4+ and CD8+

GABA accumulation can be initiated by amyloid-related processes.174

T cells in the brain as found in PD animal models.189 Interestingly,

These results were corroborated by Brawek et al.176 who showed

Watson and colleagues190 followed the regional and temporal pattern

that in amyloid-depositing mice, astrocytes accumulate around senile

of microglial activation production in mice overexpressing wild-type

type

mice.174,175

phenotype.69

plaque, increased in soma size, and produce higher levels of GABA not

human 𝛼-synuclein. They showed that after 𝛼-synuclein overexpres-

only in DG but also in frontal cortex, suggesting that altered GABA

sion, microglia were activated in the striatum already at 1 month of

production could be a general astrocyte dysfunction in AD. Interest-

age and only after 5 months in the substantia nigra, but not in the cere-

ingly, it has been shown that brain tissue from aged individuals and

bral cortex or cerebellum. The pro-inflammatory cytokine production

p16INK4a

and MMP-1

was observed only after 5–6 months of age in the brain region where

both markers of the SASP phenotype together with increased produc-

microglia was found initially activated, indicating that 𝛼-synuclein

tion of pro-inflammatory cytokines, such as IL-6.177 Moreover, other

overexpression causes a selective early inflammatory response that

authors have demonstrated that A𝛽 can act as an NF𝜅B activator lead-

was exacerbated in aged subjects together with the overexpression

ing to the release of the C3 complement component in a mouse model

of TLR4, TLR6, and TLR2 (until 14 months of age) and the recruit-

of AD. The C3 component can bind C3a receptor on neurons, thus influ-

ment of CD4- and CD8-positive cells.190 It has been also shown that

encing dendritic morphology and cognitive functions.178

TLR2 binds directly to fibrillary 𝛼-synuclein triggering TNF and IL-

patients with AD exhibit higher expression of

Finally, the glymphatic-mediated clearance of A𝛽 peptide is highly

1𝛽 production,191 whereas TLR4 by interacting with 𝛼-synuclein can

relevant in AD since emerging results underlined that age-related

mediate its uptake promoting a pro-inflammatory status character-

cerebral lymphatic vasculature shows reduced drainage of intersti-

ized by cytokine production and ROS generation in both microglia and

tial fluid and solutes, possibly leading to failure of elimination of A𝛽

astroglia.192,193 Soluble fibrillar 𝛼-synuclein recognized by cell surface

oligomers and promoting A𝛽 deposition in the brain.179 Accordingly,

receptors such as TLR2 and TLR4 can trigger NF𝜅B-dependent pro-

recent results reported in a mouse model of AD that glymphatic trans-

inflammatory gene expression and up-regulates NLRP3 component of

port was reduced, and this led to the accumulation of toxic A𝛽 soluble

the inflammasome, a cytosolic signaling complex required to transform

oligomers,180 suggesting that restoring glymphatic functionality could

inactive IL-1𝛽 and IL-18 in their active forms, thus further promoting

be a therapeutic target to slow the onset of AD.

neuroinflammation.194 Interestingly, the inflammatory process seems

PD is the second most common neurodegenerative disease after

to be regulated by small microRNAs, as in a mouse model of PD pro-

AD resulting of aging. Clinically, it is characterized by resting tremor,

duced by adeno-associated virus-mediated expression of 𝛼-synuclein,

518

SANTORO ET AL .

the loss of miR-155 reduced pro-inflammatory responses including

between immune cells (especially innate immune cells) and senescence

cytokine release, nitric oxide synthase production, and MHCII expres-

might be the common denominator in its different faces (Figure 1).

sion, thus blocking

neurodegeneration.195

This is in agreement with

During brain development, a large number of immune molecules

other reported data suggesting that the inhibition of IFN-𝛾 and TNF-

belonging to SASP seem to be involved in several aspects of “building

𝛼 production by microglia and astrocytes could be used as therapeu-

brain,” such as neuronal and glial differentiation, synaptic maturation,

tic strategies to delay neuronal degeneration in PD.196,197 As in the

and vasculogenesis through the important contribution of microglia

case of AD, the production of the above-reported pro-inflammatory

and astrocytes. Indeed, senescence and the related SASP response

cytokines are indicative of a SASP phenotype of microglia and astro-

is not a singular state and its final outcome can be influenced by

cytes that could contribute to PD progression. In this context, it has

several factors. Many of them are the peculiar molecular and cellular

been reported that environmental stressors including pesticide expo-

sensors of the tissue microenvironment triggering senescence both

sure and ROS induction caused by glutathione depletion activated

in the presence of external stressors inducing DNA damage and ROS

SASP-associated inflammatory pathways (NF𝜅B and p38MAPK) and

production and in the context of development and aging with the aim

stimulated secretion of the SASP-associated cytokine IL-6 enhancing

to promote tissue remodeling and repair. In the pathological context, it

PD risk factors.167

would be necessary to understand if the senescence that is observed in

Many other cellular processes not fully explored could be impli-

brain cells during the various neuropathologies is part of their etiology

cated in promoting neurodegeneration, especially macroautophagy.

and supports their progression, or their appearance is a consequence

From this point of view, the emerging literature demonstrates that

of the same disease. From a molecular point of view, DNA and stress

the co-chaperone and anti-apoptotic BAG3 protein,198 expressed

molecules sensing mechanisms, which regulate SASP, might have a

in glial cells and neurons during brain development and neu-

putative role. Senescence induction might be triggered first as pro-

ronal

differentiation,199

could enhance the ability of astrocytes

tective response to inner and environmental stressors encountered

to clear misfolded and aggregate proteins released from neurons,

during the life, but if sustained over the time or not properly regulated,

contributing to maintain tissue homeostasis probably also partic-

it might favor the insurgence of disease and other adverse effects of

ipating to the cytoskeletal remodeling that astrocytes undergo

senescence in vivo. Microglia might represent an essential cellular

during astrogliosis.200

component with the capacity to oscillate among some physiological

Although it is clear that glia activation is a key event in neurodegen-

and pathological conditions. It regulates neuronal plasticity in the

eration, it is still debated if innate immunity activation precedes or is

brain during development and by buffering neurotransmitters and

the consequence of neuronal death in both PD and AD201 and what

ions, modulates local blood flow, thus contributing to the permeability

is the exact role of adaptive immunity in neurodegeneration during

of BBB. This, together with SASP-specific chemokines produced also

aging. However, from the overall findings, it appears that the immune

by other patrolling brain cells like astrocytes and following senes-

response to protein accumulation triggers deleterious events such as

cence of endothelial cells, could in turn recruit immune cells from the

oxidative stress and cytokine receptor-mediated cell death, which lead

periphery, functioning as the main orchestrator of cell communication

to neuronal loss. As the inflammatory tissue milieu is the fil rouge

in the brain. The fine tune and the reciprocal influence of different

of all these aging-related disorders, an imbalance in DNA and stress

kind of cells and mediators, required to initiate senescence and its

molecules sensing mechanisms, which regulate SASP,6 might also be

spreading to neighboring cells, explain why any alteration of this

responsible for. Indeed, it is now commonly accepted that the acti-

sickly equilibrium can lead to the appearance of those conditions in

vation of glia can be neuroprotective in the first stage of the disease

which senescence has known detrimental such as neuroinflammation,

but the chronic innate immune activation could lead to a closed cir-

neurodegeneration, and glioma onset. We have also to stress that even

cuit of autosustaining inflammation involving also T-cell infiltration

though the basal levels of pro-inflammatory cytokines and chemokines

from the periphery that can favor disease progression. Future studies

are higher than in young brain, the increase in pro-inflammatory

addressing if and how pathways of senescence induction and innate

mediators in aging brain, that we could define the “brain SASP pheno-

immune responses are intimately connected will allow the identifica-

type,” are not associated necessarily with pathologies; on the contrary,

tion of novel target of pharmacological interventions to ameliorate

they serve to resolve inflammation processes by self-regulating

age-related brain disease.

and eliminating pathogens and aggregation of detrimental proteins. In this context, we believe that the time passed since senescence initiation and its paracrine effects on tissue milieu is another clue influencing the balance in senescence response; when the immune

5 CONCLUSIONS AND FUTURE DIRECTIONS

response is exacerbated by the presence of excessive stimulation, protective inflammaging may shift toward a detrimental process of neuroinflammation favoring neurodegeneration and cancer. Indeed,

As we have summarized in this paper, compelling evidence exists that

senescence features of innate immunity, in terms of morphological

the role of senescence is no longer restricted to the context of stress

changes and chronic subinflammatory status, play major role in AD

and cellular damage. In peripheral tissues and partly in brain, it occurs

and PD, however further studies are needed to find the threshold

in regulating both physiological and pathological conditions. Even

of proinflammatory status over which the disease ensues and then

though its occurrence in CNS begins to be elucidated, the interplay

innate immunity promotes rather than limits the disease. Moreover,

519

SANTORO ET AL .

& NT E PM IS LO NES E V E DE OG E U AN SS RG I O T

TU

NI N RE G O SP F ON IMM SE UN S E

↑ Influx of PBMC via BBB distruption upon EC senescence

REP A RES IR AND OLU F TIO IBROS N IS

UE

↑ pro-migratory MMPs ↓ Brain Tissue Structure and other SASp factors by for Immunesenescence of gloma-activated microglia microglia and astrocytes ↑ putative clearance of senescent pre-cancerous cells by Microglia and NK

TISS

N SIO EVA M UNE IS IMM CHAN ME

↑ Inflammaging ↑ Neurogenesis due to Senescence of NPCs ↑ Neuroinflammation Neurotrophic Role & ↑ Aged M2 ↑ Vasculogenesis microglia by SASP ↑ Clearance of aberrant proteins ↑ T-cell suppression by senescent GSCs ↓ Myelination due to OD senescence

IMMUNESURVEILLANCE F I G U R E 1 The effects of cellular senescence in different biological processes mediated by the innate immune system in the brain are schematically summarized. Conditions in which senescence are known beneficial (indicated in green) or detrimental (indicated in red) are listed. SASP, senescence-associated secretory phenotype; NPCs, neuronal progenitor cells; OD, oligodendrocytes; BBB, blood-brain barrier; GSCs, glioma stem cells; MMPs, matrix metalloproteinases

more extensive studies have to be performed to better clarify the

data; A.A.P. and E.C. wrote the paper, critically discussed literature,

role of astrocytes dysfunction in AD pathogenesis and progression

and coordinated the research. All authors approved the final version to

and to establish whether astrocyte senescence in AD precedes or

be published.

follows A𝛽 deposition. Further given the importance of the glymphatic system and the meningeal lymphatic vessels as part of a bidirectional

ACKNOWLEDGMENTS

transporter system of solutes and immune cells outside the brain toward deep cervical lymph nodes and inside the brain through perivascular/meningeal pathways, major efforts are needed to verify the senescence occurrence as molecular driver during their generation and lifespan. On the other hand, the deleterious effect of aging and the hypo- or hyperfunction of this connection in neuroinflammatory

This study was supported by University of Salerno grant ORSA 177721 to A. Santoro. E. Ciaglia was supported by a fellowship from Fondazione Umberto Veronesi (FUV 2017, cod.1072 & FUV 2018, cod.2153) (Ministry of Health (ricerca corrente) and PRIN-20157ATSLF_009 to A.A.P.).

and neurodegenerative conditions (especially those associated with protein accumulation) should be also addressed by future researches. In conclusion, understanding the biological and molecular bases of

DISCLOSURES The authors declare no conflicts of interest.

senescence, and the interplay between cellular senescence and innate immunity, which controls distinct functions in healthy and disease brain, is a challenge and an opportunity that has a clinical importance and can lead to identify new pharmacological targets to maintain or restore, when dysregulated, the physiological functions in long-lived individuals.

AUTHORSHIP All authors checked literature data articles and reviews. A.S., C.S., S.M., S.L.N., and M.C. wrote the paper and critically discussed literature

REFERENCES 1. Muñoz-Espín D, Serrano M. Cellular senescence: from physiology to pathology. Nat Rev Mol Cell Biol. 2014;15:482–496. 2. Coppe JP, Patil CK, Rodier F, et al. A human-like senescenceassociated secretory phenotype is conserved in mouse cells dependent on physiological oxygen. PLoS One. 2010;5:e9188. 3. Benarroch EE. Neuron-astrocyte interactions: partnership for normal function and disease in the central nervous system. Mayo Clin Proc. 2005;80:1326–1338.

520

4. Maciel-Barón L, Moreno-Blas D, Morales-Rosales SL, et al. Cellular senescence, neurological function, and redox state. Antioxid Redox Signal. 2017. https://doi.org/10.1089/ars.2017.7112. 5. Ransohoff RM, Brown MA. Innate immunity in the central nervous system. J Clin Invest. 2012;122:1164–1171. 6. Glück S, Guey B, Gulen MF, et al. Innate immune sensing of cytosolic chromatin fragments through cGAS promotes senescence. Nat Cell Biol. 2017;19:1061–1070. 7. Storer M, Mas A, Robert-Moreno A, et al. Senescence is a developmental mechanism that contributes to embryonic growth and patterning. Cell. 2013;155:1119–1130. 8. Krizhanovsky V, Xue W, Zender L, Yon M, Hernando E, Lowe SW. Implications of cellular senescence in tissue damage response, tumor suppression, and stem cell biology. Cold Spring Harb Symp Quant Biol. 2008;73:513–522. 9. Lampron A, Elali A, Rivest S. Innate immunity in the CNS: redefining the relationship between the CNS and Its environment. Neuron. 2013;78:214–232. 10. GBD 2015 Neurological Disorders Collaborator Group. 2017. Global, regional, and national burden of neurological disorders during 1990– 2015: a systematic analysis for the Global Burden of Disease Study 2015. Lancet Neurol. 16:877–897.

SANTORO ET AL .

26. Besancenot R, Chaligné R, Tonetti C, et al. A senescence-like cellcycle arrest occurs during megakaryocytic maturation: implications for physiological and pathological megakaryocytic proliferation. PLoS Biol. 2010;8:e1000476. 27. Cudejko C, Wouters K, Fuentes L, et al. p16INK4a deficiency promotes IL-4-induced polarization and inhibits proinflammatory signaling in macrophages. Blood. 2011;118:2556–2566. 28. Fuentes L, Wouters K, Hannou SA, et al. Downregulation of the tumour suppressor p16INK4A contributes to the polarisation of human macrophages toward an adipose tissue macrophage (ATM)like phenotype. Diabetologia. 2011;54:3150–3156. 29. Ruckh JM, Zhao JW, Shadrach JL, et al. Rejuvenation of regeneration in the aging central nervous system. Cell Stem Cell. 2012;10:96–103. 30. Kelly J, Ali Khan A, Yin J, Ferguson TA, Apte RS. Senescence regulates macrophage activation and angiogenic fate at sites of tissue injury in mice. J Clin Invest. 2007;117:3421–3426. 31. Giardino Torchia ML, Ciaglia E, Masci AM, et al. Dendritic cells/natural killer cross-talk: a novel target for human immunodeficiency virus type-1 protease inhibitors. PLoS One. 2010;5:e11052. 32. Ishikawa K, Toyama-Sorimachi N, Nakada K, et al. The innate immune system in host mice targets cells with allogenic mitochondrial DNA. J Exp Med. 2010;207:2297–2305.

11. Hayflick L. The limited in vitro lifetime of human diploid cell strains Exp. Cell Res. 1965;37:614–636.

33. Schon EA, DiMauro S, Hirano M. Human mitochondrial DNA: roles of inherited and somatic mutations. Nat Rev Genet. 2012;13:878–890.

12. Bodnar AG, Ouellette M, Frolkis M, et al. Extension of life-span by introduction of telomerase into normal human cells. Science. 1998;279:349–352.

34. Kang TW, Yevsa T, Woller N, et al. Senescence surveillance of pre-malignant hepatocytes limits liver cancer development. Nature. 2011;479:547–551.

13. Sherr CJ, DePinho RA. Cellular senescence: mitotic clock or culture shock. Cell. 2000;102:407–410.

35. Sagiv A, Biran A, Yon M, Simon J, Lowe SW, Krizhanovsky V. Granule exocytosis mediates immune surveillance of senescent cells. Oncogene. 2013;32:1971–1977.

14. von Zglinicki T. Oxidative stress shortens telomeres. Trends Biochem Sci. 2002;27:339–344. 15. Nardella C, Clohessy JG, Alimonti A, Pandolfi PP. Pro-senescence therapy for cancer treatment. Nat Rev Cancer. 2011;11:503–511. 16. Kim WY, Sharpless NE. The regulation of INK4/ ARF in cancer and aging. Cell. 2006;127:265–275. 17. Di Micco R, Cicalese A, Fumagalli M, et al. DNA damage response activation in mouse embryonic fibroblasts undergoing replicative senescence and following spontaneous immortalization. Cell Cycle. 2008;7:3601–3606. 18. Macip S, Igarashi M, Fang L, et al. Inhibition of p21-mediated ROS accumulation can rescue p21-induced senescence. EMBO J. 2002;21:2180–2188. 19. He S, Sharpless NE. Senescence in health and disease. Cell. 2017;169:1000–1011.

36. Iannello A, Thompson TW, Ardolino M, Lowe SW, Raulet DH. p53dependent chemokine production by senescent tumor cells supports NKG2D-dependent tumor elimination by natural killer cells. J Exp Med. 2013;210:2057–2069. 37. Soriani A, Zingoni A, Cerboni C, et al. ATM-ATR-dependent upregulation of DNAM-1 and NKG2D ligands on multiple myeloma cells by therapeutic agents results in enhanced NK-cell susceptibility and is associated with a senescent phenotype. Blood. 2009;113: 3503–3511. 38. Ciaglia E, Laezza C, Abate M, et al. Recognition by natural killer cells of N6-isopentenyladenosine-treated human glioma cell lines. Int J Cancer. 2018;142(1):176–190. 39. Lujambio A, Akkari L, Simon J, et al. Non-cell-autonomous tumor suppression by p53. Cell. 2013;153:449–460.

20. Sharma A, Singh K, Almasan A. Histone H2AX phosphorylation: a marker for DNA damage. Methods Mol Biol. 2012;920:613–626.

40. Egashira M, Hirota Y, Shimizu-Hirota R, et al. F4/80+ macrophages contribute to clearance of senescent cells in the mouse postpartum uterus. Endocrinology. 2017;158:2344–2353.

21. Narita M, Nunez S, Heard E, et al. Rb-mediated heterochromatin for˜ mation and silencing of E2F target genes during cellular senescence. Cell. 2003;113:703–716.

41. Porter KR, McCarthy BJ, Freels S, Kim Y, Davis FG. Prevalence estimates for primary brain tumors in the United States by age, gender, behavior, and histology. Neuro Oncol. 2010;12:520–527.

22. Dörr JR, Yu Y, Milanovic M, et al. Synthetic lethal metabolic targeting of cellular senescence in cancer therapy. Nature. 2013;501:421–425.

42. Kuilman T, Michaloglou C, Mooi WJ, Peeper DS. The essence of senescence. Genes Dev. 2010;24:2463–2479.

23. Muñoz-Espín D, Cañamero M, Maraver A, et al. Programmed cell senescence during mammalian embryonic development. Cell. 2013;155:1104–1118.

43. Toso A, Revandkar A, Di Mitri D, et al. Enhancing chemotherapy efficacy in Pten-deficient prostate tumors by activating the senescenceassociated antitumor immunity. Cell Rep. 2014;9:75–89.

24. Jun JI, Lau LF. The matricellular protein CCN1 induces fibroblast senescence and restricts fibrosis in cutaneous wound healing. Nat Cell Biol. 2010;12:676–685.

44. Eggert T, Wolter K, Ji J, et al. Distinct functions of senescenceassociated immune responses in liver tumor surveillance and tumor progression. Cancer Cell. 2016;30:533–547.

25. Krizhanovsky V, Yon M, Dickins RA, et al. Senescence of activated stellate cells limits liver fibrosis. Cell. 2008;134:657–667.

45. Joshua D, Suen H, Brown R, et al. The T cell in myeloma. Clin Lymphoma Myeloma Leuk. 2016;16:537–542.

SANTORO ET AL .

521

47. Alvarez-Buylla A, Lim D. For the long run: maintaining germinal niches in the adult brain. Neuron. 2004;41:683–686.

66. Sun H, Liang R, Yang B, et al. Aquaporin-4 mediates communication between astrocyte and microglia: implications of neuroinflammation in experimental Parkinson's disease. Neuroscience. 2016;317: 65–75.

48. Winner B, Kohl Z, Gage FH. Neurodegenerative disease and adult neurogenesis. Eur J Neurosci. 2011;33:1139–1151.

67. Farina C, Aloisi F, Meinl E. Astrocytes are active players in cerebral innate immunity. Trends Immunol. 2007;28:138–145.

49. Hamilton L, Joppé S, Cochard LM, Fernandes K. Aging and neurogenesis in the adult forebrain: what we have learned and where we should go from here. Eur J Neurosci. 2013;37:1978–1986.

68. Sofroniew MV, Vinters HV. Astrocytes: biology and pathology. Acta Neuropathol. 2010;119:7–35.

46. Gotz M, Huttner WB. The cell biology of neurogenesis. Nat Rev Mol Cell Biol. 2005;6:777–788.

50. Dong CM, Wang XL, Wang GM, et al. A stress-induced cellular aging model with postnatal neural stem cells. Cell Death Dis. 2014;5:e1116. 51. Schneider L, Pellegatta S, Favaro R, et al. DNA damage in mammalian neural stem cells leads to astrocytic differentiation mediated by BMP2 signaling through JAK-STAT. Stem Cell Rep. 2013;1: 123–138. 52. He N, Jin WL, Lok KH, Wang Y, Yin M, Wang ZJ. Amyloid-𝛽(142) oligomer accelerates senescence in adult hippocampal neural stem/progenitor cells via formylpeptide receptor 2. Cell Death Dis. 2013;4:e924. 53. Raabe E, Lim K, Kim J, et al. BRAF activation induces transformation and then senescence in human neural stem cells: a pilocytic astrocytoma model. Clin Cancer Res. 2011;17:3590–3599. 54. Ferron SR, Marques-Torrejon MA, Mira H, et al. Telomere shortening in neural stem cells disrupts neuronal differentiation and neuritogenesis. J Neurosci. 2009;29:14394–14407. 55. Bose R, Moors M, Tofighi R, Cascante A, Hermanson O, Ceccatelli S. Glucocorticoids induce long-lasting effects in neural stem cells resulting in senescence-related alterations. Cell Death Dis. 2010; 1:e92. 56. Molofsky A, Slutsky S, Joseph N, et al. Increasing p16INK4a expression decreases forebrain progenitors and neurogenesis during ageing. Nature. 2006;443:448–452. 57. Ajami B, Bennett JL, Krieger C, Tetzlaff W, Rossi FM. Local selfrenewal can sustain CNS microglia maintenance and function throughout adult life. Nat Neurosci. 2007;10:1538–1543. 58. Dong W, Cheng S, Huang F, et al. Mitochondrial dysfunction in longterm neuronal cultures mimics changes with aging. Med Sci Monit. 2011;4:BR91–BR96. 59. Geng YQ, Guan JT, Xu XH, Fu YC. Senescence-associated betagalactosidase activity expression in aging hippocampal neurons. Biochem Biophys Res Commun. 2010;396:866–869. 60. Jurk D, Wang C, Miwa S, et al. Postmitotic neurons develop a p21dependent senescence-like phenotype driven by a DNA damage response. Aging Cell. 2012;11:996–1004. 61. Bartzokis G, Lu P, Mintz J. Quantifying age-related myelin breakdown with MRI: novel therapeutic targets for preventing cognitive decline and Alzheimer's disease. J Alzheimers Dis. 2004(6Suppl):S53–S59. 62. Sim FJ, Zhao C, Penderis J, Franklin RJ. The age-related decrease in CNS remyelination efficiency is attributable to an impairment of both oligodendrocyte progenitor recruitment and differentiation. J Neurosci. 2002;22:2451–2459. 63. Tang DG, Tokumoto YM, Apperly JA, Lloyd AC, Raff MC. Lack of replicative senescence in cultured rat oligodendrocyte precursor cells. Science. 2001;291:868–871. 64. Kujuro Y, Suzuki N, Kondo T. Esophageal cancer-related gene 4 is a secreted inducer of cell senescence expressed by aged CNS precursor cells. Proc Natl Acad Sci USA. 2010;107:8259–8264. 65. Domingues HS, Portugal CC, Socodato R, Relvas JB. Oligodendrocyte, astrocyte, and microglia crosstalk in myelin development, damage, and repair. Front Cell Dev Biol. 2016;4:71.

69. Bitto A, Sell C, Crowe E, et al. Stress-induced senescence in human and rodent astrocytes. Exp Cell Res. 2010;316:2961–2968. 70. Zhang G, Li J, Purkayastha S, et al. Hypothalamic programming of systemic ageing involving IKK-𝛽, NF-𝜅B and GnRH. Nature. 2013;497:211–216. 71. Salminen LE, Paul RH. Oxidative stress and genetic markers of suboptimal antioxidant defense in the aging brain: a theoretical review. Rev Neurosci. 2014;25:805–819. 72. Mitteldorf J. Is programmed aging a cause for optimism. Curr Aging Sci. 2015;8:69–75. 73. Santos CL, Roppa PHA, Truccolo P, et al. Age-dependent neurochemical remodeling of hypothalamic astrocytes. Mol Neurobiol. 2017. https://doi.org/10.1007/s12035-017-0786-x. 74. Porchet R, Probst A, Bouras C, Draberova E, Draber P, Riederer BM. Analysis of glial acidic fibrillary protein in the human entorhinal cortex during aging and in Alzheimer's disease. Proteomics. 2003;3:1476–1485. 75. Nichols NR, Day JR, Laping NJ, Johnson SA, Finch CE. GFAP mRNA increases with age in rat and human brain. Neurobiol Aging. 1993;14:421–429. 76. Pekny M, Pekna M. Astrocyte intermediate filaments in CNS pathologies and regeneration. J Pathol. 2004;204:428–437. 77. VanGuilder HD, Bixler GV, Brucklacher RM, et al. Concurrent hippocampal induction of MHC II pathway components and glial activation with advanced aging is not correlated with cognitive impairment. J Neuroinflammation. 2011;8:138. 78. Norden DM, Godbout JP. Review: microglia of the aged brain: primed to be activated and resistant to regulation. Neuropathol Appl Neurobiol. 2013;39:19–34. 79. Harry GJ. Microglia during development and aging. Pharmacol Ther. 2013;139:313–326. 80. Streit WJ, Sammons NW, Kuhns AJ, Sparks DL. Dystrophic microglia in the aging human brain. Glia. 2004;45:208–212. 81. Flanary BE, Streit WJ. Progressive telomere shortening occurs in cultured rat microglia, but not astrocytes. Glia. 2004;45:75–88. 82. Caldeira C, Oliveira AF, Cunha C, et al. Microglia change from a reactive to an age-like phenotype with the time in culture. Front Cell Neurosci. 2014;8:152. 83. Sheng JG, Mrak RE, Griffin WS. Enlarged and phagocytic, but not primed, interleukin-1 alphaimmunoreactive microglia increase with age in normal human brain. Acta Neuropathol. 1998;5: 229–234. 84. Sierra A, Gottfried-Blackmore AC, McEwen BS, Bulloch K. Microglia derived from aging mice exhibit an altered inflammatory profile. Glia. 2007;55:412–424. 85. Hickman SE, Kingery ND, Ohsumi TK, et al. The microglial sensome revealed by direct RNA sequencing. Nat Neurosci. 2013;16: 1896–18905. 86. Salvi V, Sozio F, Sozzani S, Del Prete A. Role of atypical chemokine receptors in microglial activation and polarization. Front Aging Neurosci. 2017;9:148.

522

SANTORO ET AL .

87. Dilger RN, Johnson RW. Aging, microglial cell priming, and the discordant central inflammatory response to signals from the peripheral immune system. J Leukoc Biol. 2008;84:932–939.

106. Schläger C, Körner H, Krueger M, et al. Effector T-cell trafficking between the leptomeninges and the cerebrospinal fluid. Nature. 2016;530:349–353.

88. Henry CJ, Huang Y, Wynne AM, Godbout JP. Peripheral lipopolysaccharide (LPS) challenge promotes microglial hyperactivity in aged mice that is associated with exaggerated induction of both proinflammatory IL-1beta and anti-inflammatory IL-10 cytokines. Brain Behav Immun. 2009;23:309–317.

107. Sun BL, Wang LH, Yang T, et al. Lymphatic drainage system of the brain: a novel target for intervention of neurological diseases. Prog Neurobiol. 2017. https://doi.org/10.1016/j.pneurobio.2017.08.007.

89. Frank MG, Barrientos RM, Biedenkapp JC, Rudy JW, Watkins LR, Maier SF. mRNA up-regulation of MHC II and pivotal proinflammatory genes in normal brain aging. Neurobiol Aging. 2006;27:717–722. 90. Franceschi C, Bonafè M, Valensin S, et al. Inflamm-aging. An evolutionary perspective on immunosenescence. Ann N Y Acad Sci. 2000;908:244–254. 91. Franceschi C, Capri M, Monti D, et al. Inflammaging and antiinflammaging: a systemic perspective on aging and longevity emerged from studies in humans. Mech Ageing Dev. 2007;128: 92–105. 92. Qi Q, Liu Y, Cheng Y, et al. Diversity and clonal selection in the human T-cell repertoire. Proc Natl Acad Sci USA. 2014;111:13139–13144. 93. Qi Q, Zhang DW, Weyand CM, Goronzy JJ. Mechanisms shaping thenaive T cell repertoire in the elderly—thymic involution or peripheralhomeostatic proliferation. Exp Gerontol. 2014;54:71–74. 94. Di Benedetto S, Müller L, Wenger E, Düzel S, Pawelec G. Contribution of neuroinflammation and immunity to brain aging and the mitigating effects of physical and cognitive interventions. Neurosci Biobehav Rev. 2017;75:114–128.

108. Iliff JJ, Wang M, Zeppenfeld DM, et al. Cerebral arterial pulsation drives paravascular CSF-interstitial fluid exchange in the murine brain. J Neurosci. 2013;33:18190–18199. 109. Nedergaard M. Neuroscience. Garbage truck of the brain. Science. 2013;340:1529–1530. 110. Louveau A, Smirnov I, Keyes TJ, et al. Structural and functional features of central nervous system lymphatic vessels. Nature. 2015;523:337–341. 111. Xie L, Kang H, Xu Q, et al. Sleep drives metabolite clearance from the adult brain. Science. 2013;342:373–377. 112. Aspelund A, Antila S, Proulx ST, et al. A dural lymphatic vascular system that drains brain interstitial fluid and macromolecules. J Exp Med. 2015;212:991–999. 113. Card CM, Yu SS, Swartz MA. Emerging roles of lymphatic endothelium in regulating adaptive immunity. J Clin Invest. 2014;124: 943–952. 114. Randolph GJ, Ivanov S, Zinselmeyer BH, Scallan JP. The lymphatic system: integral roles in immunity. Annu Rev Immunol. 2017;35: 31–52. 115. Ma Q, Ineichen BV, Detmar M, Proulx ST. Outflow of cerebrospinal fluid is predominantly through lymphatic vessels and is reduced in aged mice. Nat Commun. 2017;8:1434.

95. Lana D, Iovino L, Nosi D, Wenk GL, Giovannini MG. The neuronastrocyte-microglia triad involvement in neuroinflammaging mechanisms in the CA3 hippocampus of memory-impaired aged rats. Exp Gerontol. 2016;83:71–88.

116. Petrova TV, Koh GY. Organ-specific lymphatic vasculature: from development to pathophysiology. J Exp Med. 2018;215: 35–49.

96. Kabba JA, Xu Y, Christian H, et al. Microglia: housekeeper of the central nervous system. Cell Mol Neurobiol. 2017;38:53–71.

117. Pousset F. Developmental expression of cytokine genes in the cortex and hippocampus of the rat central nervous system. Brain Res Dev Brain Res. 1994;81:143–146.

97. Gemma C, Bachstetter AD, Bickford PC. Neuron-microglia dialogue and hippocampal neurogenesis in the aged brain. Aging Dis. 2010;1:232–244.

118. Schmitz T, Chew LJ. Cytokines and myelination in the central nervous system. Scientific World J. 2008;8:1119–1147.

98. Wu Y, Dissing-Olesen L, MacVicar BA, Stevens B. Microglia: dynamic mediators of synapse development and plasticity. Trends Immunol. 2015;36:605–613.

119. Alfonsi F, Filippi P, Salaun D, deLapeyrie‘re O, Durbec P. LIFR beta plays a major role in neuronal identity determination and glial differentiation in the mouse facial nucleus. Dev Biol. 2008;313: 267–278.

99. Nicholas RS, Stevens S, Wing MG, Compston DA. Microglia derived IGF-2 prevents TNF𝛼 induced death of mature oligodendrocytes in vitro. J Neuroimmunol. 2002;124:36–44. 100. Barak B, Feldman N, Okun E. Toll-like receptors as developmental tools that regulate neurogenesis during development: an update. Front Neurosci. 2014;8:272. 101. Schmitt CA. The persistent dynamic secrets of senescence. Nat Cell Biol. 2016;18:913–915. 102. Mikuła-Pietrasik J, Sosińska P, Janus J, et al. Bystander senescence in human peritoneal mesothelium and fibroblasts is related to thrombospondin-1-dependent activation of transforming growth factor-𝛽1. Int J Biochem Cell Biol. 2013;45:2087–2096. 103. Fletcher JM, Lalor SJ, Sweeney CM, Tubridy N, Mills KHG. T cells in multiple sclerosis and experimental autoimmune encephalomyelitis. Clin Exp Immunol. 2010;162:1–11. 104. Corraliza I. Recruiting specialized macrophages across the borders to restore brain functions. Front Cell Neurosci. 2014;8:262. 105. Louveau A, Plog BA, Antila S, Alitalo K, Nedergaard M, Kipnis J. Understanding the functions and relationships of the glymphatic system and meningeal lymphatics. J Clin Invest. 2017;127:3210–3219.

120. Bachiller D, Klingensmith J, Kemp C, et al. The organizer factors Chordin and Noggin are required for mouse forebrain development. Nature. 2000;403:658–661. 121. Giulian D, Woodward J, Young DG, Krebs JF, Lachman LB. Interleukin-1 injected into mammalian brain stimulates astrogliosis and neovascularization. J Neurosci. 1988;8:2485–2490. 122. Schwarz JM, Sholar PW, Bilbo SD. Sex differences in microglial colonization of the developing rat brain. J Neurochem. 2012;120: 948–963. 123. Fee D, Grzybicki D, Dobbs M, et al. Interleukin 6 promotes vasculogenesis of murine brain microvessel endothelial cells. Cytokine. 2000;12:655–665. 124. Gadient RA, Otten U. Identification of interleukin-6 (IL-6)-expressing neurons in the cerebellum and hippocampus of normal adult rats. Neurosci Lett. 1994;182:243–246. 125. Harry GJ, Kraft AD. Microglia in the developing brain: a potential target with lifetime effects. Neurotoxicology. 2012;33:191–206. 126. Wang CC, Wu CH, Shieh JY, Wen CY. Microglial distribution and apoptosis in fetal rat brain. Brain Res Dev. 2002;139:337–342.

SANTORO ET AL .

523

127. Rice D, Barone S. Critical periods of vulnerability for the developing nervous system: evidence from humans and animal models. Environ Health Perspect. 2000;108:511–533.

150. Abate M, Laezza C, Pisanti S, et al. Deregulated expression and activity of farnesyl diphosphate synthase (FDPS) in glioblastoma. Sci Rep. 2017;7:14123.

128. Davoust N, Vuaillat C, Androdias G, Nataf S. From bone marrow to microglia: barriers and avenues. Trends Immunol. 2008;29:227–234.

151. Stoll EA, Horner PJ, Rostomily RC. The impact of age on oncogenic potential: tumor-initiating cells and the brain microenvironment. Aging Cell. 2013;12:733–741.

129. Ginhoux F, Greter M, Leboeuf M, et al. Fate mapping analysis reveals that adult microglia derive from primitive macrophages. Science. 2010;330:841–845. 130. Monier A, Adle-Biassette H, Delezoide AL, Evrard P, Gressens P, Verney C. Entry and distribution of microglial cells in human embryonic and fetal cerebral cortex. J Neuropathol Exp Neurol. 2007;66:372–382. 131. Pont-Lezica L, Béchade C, Belarif-Cantaut Y, Pascual O, Bessis A. Physiological roles of microglia during development. J Neurochem. 2011;119:901–908. 132. Antony JM, Paquin A, Nutt SL, Kaplan DR, Miller FD. Endogenous microglia regulate development of embryonic cortical precursor cells. J Neurosci Res. 2011;89:286–298. 133. Deverman BE, Patterson PH. Cytokines and CNS development. Neuron. 2009;64:61–78. 134. Deng W, Pleasure J, Pleasure D. Progress in periventricular leukomalacia. Arch Neurol. 2008;65:1291–1295. 135. Skoff RP. Gliogenesis in rat optic nerve: astrocytes are generated in a single wave before oligodendrocytes. Dev Biol. 1990;139:149–168. 136. Yang Y, Higashimori H, Morel L. Developmental maturation of astrocytes and pathogenesis of neurodevelopmental disorders. J Neurodev Disord. 2013;5:22. 137. Roessmann U, Gambetti P. Astrocytes in the developing human brain. An immunohistochemical study. Acta Neuropathol. 1986;70:308–313. 138. Ma S, Kwon HJ, Huang Z. A functional requirement for astroglia in promoting blood vessel development in the early postnatal brain. PLoS One. 2012;7:e48001. 139. Stone J, Itin A, Alon T, et al. Development of retinal vasculature is mediated by hypoxia-induced vascular endothelial growth factor (VEGF) expression by neuroglia. J Neurosci. 1995;15:4738–4747. 140. Scott A, Powner MB, Gandhi P, et al. Astrocyte-derived vascular endothelial growth factor stabilizes vessels in the developing retinal vasculature. PLoS One. 2010;5:e11863. 141. Abbott NJ, Rönnbäck L, Hansson E. Astrocyte-endothelial interactions at the blood-brain barrier. Nat Rev Neurosci. 2006;7:41–53. 142. Alvarez JI, Katayama T, Prat A. Glial influence on the blood brain barrier. Glia. 2013;61:1939–1958. 143. Lee H, Pienaar IS. Disruption of the blood-brain barrier in Parkinson's disease: curse or route to a cure. Front Biosci. 2014;19:272–280. 144. Shaw AC, Goldstein DR, Montgomery RR. Age-dependent dysregulation of innate immunity. Nat Rev Immunol. 2013;13:875–887. 145. Lalancette-Hebert M, Gowing G, Simard A, Weng YC, Kriz J. Selective ablation of proliferating microglial cells exacerbates ischemic injury in the brain. J Neurosci. 2007;27:2596–2605. 146. Baker DJ, Wijshake T, Tchkonia T, et al. Clearance of p16Ink4apositive senescent cells delays ageing-associated disorders. Nature. 2011;479:232–236. 147. Zhu Y, Tchkonia T, Pirtskhalava T, et al. The Achilles’ heel of senescent cells: from transcriptome to senolytic drugs. Aging Cell. 2015;14: 644–658.

152. Shete S, Hosking FJ, Robertson LB, et al. Genome-wide association study identifies five susceptibility loci for glioma. Nat Genet. 2009;41:899–904. 153. Walsh KM, Codd V, Smirnov IV, et al. Variants near TERT and TERC influencing telomere length are associated with high-grade glioma risk. Nat Genet. 2014;46:731–735. 154. Wheeler CJ, Black KL, Liu G, et al. Thymic CD8+ T cell production strongly influences tumor antigen recognition and age-dependent glioma mortality. J Immunol. 2003;171:4927–4933. 155. Held-Feindt J, Hattermann K, Muerkoster SS, et al. CX3CR1 promotes recruitment of human glioma-infiltrating microglia/ macrophages (GIMs). Exp Cell Res. 2010;316:1553–1566. 156. Domenis R, Cesselli D, Toffoletto B, et al. Systemic T cells immunosuppression of glioma stem cell-derived exosomes is mediated by monocytic myeloid-derived suppressor cells. PLoS One. 2017; 12:e0169932. 157. Streit WJ, Xue QS. Human CNS immune senescence and neurodegeneration. Curr Opin Immunol. 2014;29:93–96. 158. Coder B, Wang W, Wang L, Wu Z, Zhuge Q, Su DM. Friend or foe: the dichotomous impact of T cells on neuro-de/re-generation during aging. Oncotarget. 2017;8:7116–7137. 159. Molteni M, Rossetti C. Neurodegenerative diseases: the immunological perspective. J Neuroimmunol. 2017;313:109–115. 160. Bamberger ME, Harris ME, McDonald DR, Husemann J, Landreth GE. A cell surface receptor complex for fibrillar betaamyloid mediates microglial activation. J Neurosci. 2003;23: 2665–2674. 161. Fassbender K, Walter S, Kühl S, et al. The LPS receptor (CD14) links innate immunity with Alzheimer's disease. FASEB J. 2004;18: 203–205. 162. Tahara K, Kim HD, Jin JJ, Maxwell JA, Li L, Fukuchi K. Role of toll-like receptor signalling in Abeta uptake and clearance. Brain. 2006;129:3006–3019. 163. Stewart CR, Stuart LM, Wilkinson K, et al. CD36 ligands promote sterile inflammation through assembly of a Toll-like receptor 4 and 6 heterodimer. Nat Immunol. 2010;11:155–161. 164. Montgomery SL, Mastrangelo MA, Habib D, et al. Ablation of TNFRI/RII expression in Alzheimer's disease mice leads to an unexpected enhancement of pathology: implications for chronic pan-TNF𝛼 suppressive therapeutic strategies in the brain. Am J Pathol. 2011; 179:2053–2070. 165. Koenigsknecht J, Landreth G. Microglial phagocytosis of fibrillar beta-amyloid through a beta1 integrin-dependent mechanism. J Neurosci. 2004;24:9838–9846. 166. Reed-Geaghan EG, Savage JC, Hise AG, Landreth GE. CD14 and toll-like receptors 2 and 4 are required for fibrillar A{beta}-stimulated microglial activation. J Neurosci. 2009;29: 11982–11992.

148. Bilbo S, Schwarz JM. Early-life programming of later-life brain and behavior: a critical role for the immune system. Front Behav Neurosci. 2009;3:14.

167. Lee M, Cho T, Jantaratnotai N, Wang YT, McGeer E, McGeer PL. Depletion of GSH in glial cells induces neurotoxicity: relevance to aging and degenerative neurological diseases. FASEB J. 2010;24:2533–2545.

149. Vilhardt F. Microglia: phagocyte and glia cell. Int. J Biochem Cell Biol. 2005;37:17–21.

168. Dheen ST, Kaur C, Ling EA. Microglial activation and its implications in the brain diseases. Curr Med Chem. 2007;14:1189–1197.

524

169. Glass CK, Saijo K, Winner B, Marchetto MC, Gage FH. Mechanisms underlying inflammation in neurodegeneration. Cell. 2010;140: 918–934. 170. Ramasamy R, Yan SF, Schmidt AM. RAGE: therapeutic target and biomarker of the inflammatory response–the evidence mounts. J Leukoc Biol. 2009;86:505–512. 171. Bu G. Apolipoprotein E and its receptors in Alzheimer's disease: pathways, pathogenesis and therapy. Nat Rev Neurosci. 2009;10: 333–344. 172. Holtzman DM, Herz J, Bu G. Apolipoprotein E and apolipoprotein E receptors: normal biology and roles in Alzheimer disease. Cold Spring Harb Perspect Med. 2012;2:a006312. 173. Osborn LM, Kamphuis W, Wadman WJ, Hol EM. Astrogliosis: an integral player in the pathogenesis of Alzheimer's disease. Prog Neurobiol. 2016;144:121–141. 174. Jo S, Yarishkin O, Hwang YJ, et al. GABA from reactive astrocytes impairs memory in mouse models of Alzheimer's disease. Nat Med. 2014;20:886–896. 175. Wu Z, Guo Z, Gearing M, Chen G. Tonic inhibition in dentate gyrus impairs long-term potentiation and memory in an Alzhiemer's disease model. Nat Commun. 2014;5:4159. 176. Brawek B, Chesters R, Klement D, et al. A bell-shaped dependence between amyloidosis and GABA accumulation in astrocytes in a mouse model of Alzheimer's disease. Neurobiol Aging. 2018;61: 187–197. 177. Bhat R, Crowe EP, Bitto A, et al. Astrocyte senescence as a component of Alzheimer's disease. PLoS One. 2012;7:e45069. 178. Lian H, Litvinchuk A, Chiang AC, Aithmitti N, Jankowsky JL, Zheng H. Astrocyte-microglia cross talk through complement activation modulates amyloid pathology in mouse models of Alzheimer's disease. J Neurosci. 2016;36:577–589. 179. Louveau A, Da Mesquita S, Kipnis J. Lymphatics in neurological disorders: a neuro-lympho-vascular component of multiple sclerosis and Alzheimer's disease. Neuron. 2016;91:957–973.

SANTORO ET AL .

187. Blum-Degen D, Müller T, Kuhn W, Gerlach M, Przuntek H, Riederer P. Interleukin-1 beta and interleukin-6 are elevated in the cerebrospinal fluid of Alzheimer's and de novo Parkinson's disease patients. Neurosci Lett. 1995;202:17–20. 188. Dufek M, Rektorova I, Thon V, Lokaj J, Rektor I. Interleukin-6 may contribute to mortality in Parkinson's disease patients: a 4-year prospective study. Parkinsons Dis. 2015;2015:898192. 189. Brochard V, Combadière B, Prigent A, et al. Infiltration of CD4+ lymphocytes into the brain contributes to neurodegeneration in a mouse model of Parkinson disease. J Clin Invest. 2009;119: 182–192. 190. Watson MB, Richter F, Lee SK, et al. Regionally-specific microglial activation in young mice over-expressing human wildtype alphasynuclein. Exp Neurol. 2012;237:318–334. 191. Kim C, Ho DH, Suk JE, et al. Neuron-released oligomeric 𝛼-synuclein is an endogenous agonist of TLR2 for paracrine activation of microglia. Nat Commun. 2013;4:1562. 192. Fellner L, Irschick R, Schanda K, et al. Toll-like receptor 4 is required for 𝛼-synuclein dependent activation of microglia and astroglia. Glia. 2013;61:349–360. 193. Gustot A, Gallea JI, Sarroukh R, Celej MS, Ruysschaert JM, Raussens V. Amyloid fibrils are the molecular trigger of inflammation in Parkinson's disease. Biochem J. 2015;471:323–333. 194. Labzin LI, Heneka MT, Latz E. Innate immunity and neurodegeneration. Annu Rev Med. 2017. https://doi.org/10.1146/annurev-med050715-104343. 195. Thome AD, Harms AS, Volpicelli-Daley LA, Standaert DG. microRNA155 regulates alpha-synuclein-induced inflammatory responses in models of Parkinson disease. J Neurosci. 2016;36:2383–2390. 196. Hashioka S, Klegeris A, Schwab C, McGeer PL. Interferon-gammadependent cytotoxic activation of human astrocytes and astrocytoma cells. Neurobiol Aging. 2009;30:1924–1935.

180. Peng W, Achariyar TM, Li B, et al. Suppression of glymphatic fluid transport in a mouse model of Alzheimer's disease. Neurobiol Dis. 2016;93:215–225.

197. Neves KR, Nobre HV, Jr., Leal LK, de Andrade GM, Brito GA, Viana GS. Pentoxifylline neuroprotective effects are possibly related to its anti-inflammatory and TNF-alpha inhibitory properties, in the 6-OHDA model of Parkinson's disease. Parkinsons Dis. 2015; 2015:108179.

181. Liang Z, Zhao Y, Ruan L, et al. Impact of aging immune system on neurodegeneration and potential immunotherapies. Prog Neurobiol. 2017;157:2–28.

198. Guerriero L, Chong K, Franco R, et al. BAG3 protein expression in melanoma metastatic lymph nodes correlates with patients’ survival. Cell Death Dis. 2014;5:e1173.

182. González H, Contreras F, Prado C, et al. Dopamine receptor D3 expressed on CD4+ T cells favors neurodegeneration of dopaminergic neurons during Parkinson's disease. J Immunol. 2013;190: 5048–5056.

199. Santoro A, Nicolin V, Florenzano F, Rosati A, Capunzo M, Nori SL. BAG3 is involved in neuronal differentiation and migration. Cell Tissue Res. 2017;368:249–258.

183. Double KL, Rowe DB, Carew-Jones FM, et al. Anti-melanin antibodies are increased in sera in Parkinson's disease. Exp Neurol. 2009;217:297–301. 184. Shameli A, Xiao W, Zheng Y, et al. A critical role for alpha-synuclein in development and function of T lymphocytes. Immunobiology. 2016;221:333–340. 185. Park JY, Kim KS, Lee SB, et al. On the mechanism of internalization of alpha-synuclein into microglia: roles of ganglioside GM1 and lipid raft. J Neurochem. 2009;110:400–411. 186. Zhang W, Wang T, Pei Z, et al. Aggregated alpha-synuclein activates microglia: a process leading to disease progression in Parkinson's disease. FASEB J. 2005;19:533–542.

200. Seidel K, Vinet J, Dunnen WF, et al. The HSPB8-BAG3 chaperone complex is upregulated in astrocytes in the human brain affected by protein aggregation diseases. Neuropathol Appl Neurobiol;38: 39–53. 201. Hirsch EC, Hunot S. Neuroinflammation in Parkinson's disease: a target for neuroprotection. Lancet Neurol. 2009;8:382–397.

How to cite this article:

Santoro A, Spinelli CC, Martuc-

ciello S, et al. Innate immunity and cellular senescence: The good and the bad in the developmental and aged brain. J Leukoc Biol. 2018;103:509–524. https://doi.org/10.1002/JLB. 3MR0118-003R