Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

FOCUS | Review Article

FOCUS | Review Article https://doi.org/10.1038/s41593-018-0242-x

https://doi.org/10.1038/s41593-018-0242-x

Microglia in neurodegeneration
Suzanne Hickman, Saef Izzy   , Pritha Sen, Liza Morsett and Joseph El Khoury   *

The neuroimmune system is involved in development, normal functioning, aging, and injury of the central nervous system.
Microglia, first described a century ago, are the main neuroimmune cells and have three essential functions: a sentinel function
involved in constant sensing of changes in their environment, a housekeeping function that promotes neuronal well-being and
normal operation, and a defense function necessary for responding to such changes and providing neuroprotection. Microglia
use a defined armamentarium of genes to perform these tasks. In response to specific stimuli, or with neuroinflammation,
microglia also have the capacity to damage and kill neurons. Injury to neurons in Alzheimer’s, Parkinson’s, Huntington’s, and
prion diseases, as well as in amyotrophic lateral sclerosis, frontotemporal dementia, and chronic traumatic encephalopathy,
results from disruption of the sentinel or housekeeping functions and dysregulation of the defense function and neuroinflam-
mation. Pathways associated with such injury include several sensing and housekeeping pathways, such as the Trem2, Cx3cr1
and progranulin pathways, which act as immune checkpoints to keep the microglial inflammatory response under control, and
the scavenger receptor pathways, which promote clearance of injurious stimuli. Peripheral interference from systemic inflam-
mation or the gut microbiome can also alter progression of such injury. Initiation or exacerbation of neurodegeneration results
from an imbalance between these microglial functions; correcting such imbalance may be a potential mode for therapy.

R
ecent research into microglia provides unprecedented insight monocytes to develop into microglia-like cells is possible15–17 and is
into their roles in health, aging, and neurodegenerative dis- dependent on their environment18.
eases. These advances started 100 years ago in 1918, when Until recently, a simplistic definition of microglia describes them
Pio del Rio Hortega published a method for staining microglia and as innate immune cells of the CNS of myeloid origin that express
distinguishing them from neighboring cells of the CNS1. Hortega Cx3cr1, CD11b, Iba1, and F4/8011. Based on comprehensive gene
named microglia the ‘third element’ of the CNS, describing their expression profiling and functional studies7,11, we propose a func-
phagocytic function, plasticity, regional distribution, and het- tional and molecular definition of microglia that correlates their
erogeneity. For a century, microgliologists have been validating gene expression with their functions. RNA-seq analysis identified
Hortega’s observations. a new set of microglia-specific markers in the healthy brain that
Development of methods to isolate and culture neonatal include HexB, P2ry12, S100A8, S100A9, Tmem119, Gpr34, SiglecH,
microglia2 ascertained their functions, including phagocyto- TREM2, and Olfml37. Microglial transcriptomes allow them to
sis and response to amyloid-β​ (Aβ​), and supported their roles in perform three essential functions: (i) sense their environment,
neurodegeneration. Generation of mice with GFP-labeled microg- (ii) conduct physiological housekeeping, and (iii) protect against
lia3 allowed in vivo visualization by two-photon microscopy and modified-self and non-self injurious agents. These normal functions
showed that microglia continually survey and sense their micro- are important in various stages of development from embryonic
environment, respond rapidly to focal injury4, are involved in stages to adulthood and aging.
synaptic pruning and remodeling5, and contribute to various neu-
rodegenerative diseases. Novel methods to isolate adult microglia6 Sensing. Microglia form a network spanning the CNS9. Their thin
allowed transcriptomic analyses by RNA sequencing, thus identi- processes are dynamic and in constant motion, allowing them to
fying expression signatures that help define these cells7. Recently, scan the area surrounding their cell body every few hours and rapidly
single-cell RNA-seq has provided insight into potential microglial polarize toward focal injury (Fig. 1a–c and Supplementary Video 1).
subpopulations in neurodegenerative diseases8. They use the products of nearly 100 genes to sense changes in their
In this Review, we summarize the current knowledge of the microenvironment (their sensome) including P2yr12, AXL, and
roles of microglia in neurodegeneration. To better understand MER4,7,19 (Fig. 1d,e). Sensome mRNAs are uniformly expressed in
such roles, we introduce a revised functional and transcriptomic microglia in various areas of the brain, suggesting that all microglia
definition of microglia, discuss their roles in individual neuro- are capable of performing their sensing function (Fig. 1e). Sensing
degenerative diseases, and review common pathways involved in is a prerequisite for microglia to perform their housekeeping and
neurodegeneration. host defense functions.

A functional and molecular definition of microglia Housekeeping. Physiological housekeeping functions include
Microglia constitute 5–12% of CNS cells, depending on the region9. synaptic remodeling (a function critical for CNS development,
They are the principal resident immune cells of the brain and are homeostasis, and neurodegeneration20–22), migration to sites of neu-
involved in homeostasis and in host defense against pathogens ronal death to phagocytose dead or dying cells23,24 or debris, and
and CNS disorders10,11. Ontological studies of microglia confirmed maintaining myelin homeostasis25. Interacting with astrocytes is
Hortega’s suspicion that they are mesenchymal, myeloid12, origi- another important microglial function involved in homeostasis,
nating in the yolk sac, and capable of self-renewal independent of inflammation, and possibly neurodegeneration26. Among the genes
hematopoietic stem cells13. Microglial survival and maintenance involved in housekeeping are those encoding chemokine and che-
depend on cytokines, including CSF1 and interleukin (IL)-3414, and moattractant receptors, genes involved in phagocytosis (scavenger
on transcription factors such as IRF812. Reprograming stem cells or receptors and Trem2), and genes involved in synaptic pruning and

Center for Immunology & Inflammatory Diseases, Massachusetts General Hospital, and Harvard Medical School, Boston, MA, USA.
*e-mail: jelkhoury@mgh.harvard.edu

Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience 1359


Review Article | FOCUS
Review Article | FOCUS
https://doi.org/10.1038/s41593-018-0242-x NATure NeurosCIenCe

Microglial network spans the brain Ontology of the sensome


Cytokine receptors (10%)
a b c d
Pattern-recognition
receptors (25%) Purinergic receptors
(8%)

ECM receptors
(6%)
50 μm 50 μm 50 μm
Chemoattractant and
Cortex Hippocampus Cerebellum chemokines receptors (10%)

20 μm
e Uniform expression of sensome transcripts across the brain

Cerebellum

Medulla

Pons

Midbrain

Hypothalamus

Thalamus

Striatum

Hippocampus

Olfactory bulb

Isocortex
Tmem119

Tmem173
Tnfrsf13b

Tmem37

Tmem8c
Tnfrsf17

Tnfrsf1b

Clec4a3
Clec4a2
Slco2b1

Slc16a3
Csf2rb2
Gpr160

Gpr183
Siglech

Siglec5

Adora3

Cmtm7
P2ry12
Entpd1

P2yr13

Cx3cr1

Cmklr1
Fcgr2b

Fcer1g

Slamf9

Clec7a

Cysltr1

Clec5a
Cxcl16

Slc2a5

Slc7a7
Cd101

Lgals9

Dap12

Cd180

Cd79b
Tgfbr2

Upk1b

Tgfbr1
Selplg

Gpr84

Gpr34

Gpr77
Trem2

C5ar1

C3ar1

Icam4

Icam1
P2ry6
Fcgr1

Ifngr1

Fcgr3

Fcgr4

Lpar5
Ecscr

Itgam

Il10ra
Lilra5
Csf1r

Cd37

Emr1
Cd68

Cd74
Cd14

Cd86

Cd33

Cd48

Cd22

Cd53

Cd52

Cd84

ifitm6
Ccrl2
Ptprc
Tlr12

Tlr13

FcrI1
Lag3

Itgb5

Lair1

Itgb2

Ly86
Ccr5

Ptafr

Ccr5

Pilra
Il21r

Il6ra

Tlr6

Tlr2

Tlr1

Tlr7

Tlr4

Ltf
Fig. 1 | Microglia in a normal mouse brain. a–c, Mouse microglia, stained here with anti-CD11b, have distinct processes that are constantly moving in the
area around the cell body, and form a network of cells that spans most of the CNS, including the (a) cortex (b) hippocampus, and (c) cerebellum. d, Three-
dimensional image of a mouse microglia with summary of gene ontology analysis of the sensome genes. e, Heatmap showing comparative expression of
microglial sensome genes identified by RNA-seq data using the Allen Brain Atlas in situ hybridization dataset. Most of the genes are similarly expressed in
most areas of the brain, except for two small clusters that appear to have differential expression in the brain stem. ECM, extracellular matrix.

remodeling (C1q and Cx3cr1; Fig. 2)7. Aberrant housekeeping can Alzheimer’s disease
lead to neurodegeneration. Alzheimer’s disease (AD) is characterized by formation of Aβ​-
containing plaques, neurofibrillary tangles comprising intracel-
Protection against injurious self and non-self stimuli. Microglia lular hyperphosphorylated tau protein, and neuronal loss29. An
mediate host defense against infectious pathogens, injurious self- accepted sequence of events is that accumulation of Aβ​leads to
proteins such as Aβ​, aggregated α​-synuclein, mutant hunting- a microglial response, which promotes tau hyperphosphorylation
tin, mutant or oxidized superoxide dismutase (SOD), or prions, and formation of neurofibrillary tangles, leading to neurodegener-
as well as primary or metastatic CNS tumors. To perform these ation and cognitive impairment. In AD patients and animal mod-
functions, microglia express Fc receptors, Toll-like receptors els, microglia accumulate around senile plaques (Fig. 2c), where
(TLRs), viral receptors, and antimicrobial peptides (Fig. 2)7. In their density is two- to fivefold higher than in normal paren-
response to such stimuli, microglia can initiate a neuroinflam- chyma30. They contain intracellular Aβ​, suggesting phagocytosis31,
matory response which, like peripheral inflammation, includes show proinflammatory morphological changes such as somatic
production of cytokines such as TNF and IL-16,27, and possibly swelling and process shortening (Fig. 2c), and have increased
chemokines such as Ccl228, to recruit additional cells and induce proinflammatory markers including major histocompatibility
them to clear injurious agents and maintain brain homeostasis. complex II, CD36, IL-1, IL-6, and TNF32,33. So how do microglia
Neuroinflammation, however, unlike peripheral inflammation, contribute to AD pathogenesis?
can also be limited to microglia without recruiting circulating
leukocytes. Persistent neuroinflammation in turn induces neuro- Genome-wide association studies. Evidence for a direct microg-
toxicity, leading to neurodegeneration. lial role in AD came from genome-wide association studies.
A take-home message is that there are no resting microglia Mutations in triggering receptor expressed on myeloid cells 2
(Fig. 2). Their sensing, housekeeping, and protecting functions keep (Trem2) were associated with a 3.0- to 4.5-fold increased AD risk,
them constantly engaged, and most microglia in healthy brains are almost as high as that associated with ApoE ε​434,35. Mutations in
capable of performing such functions. Dysregulation of any of other microglial genes, such as CR1, HLA–DRB1, CD33, MS4A6A,
these functions results in an imbalance that initiates or propagates and BIN1, were associated with more modest AD risks34. Since
neurodegeneration. Here we summarize what we know about these genes regulate key microglial functions, understanding how
microglia and what happens to their functions in various neurode- they affect AD will impact all AD patients whether they have these
generative diseases. mutations or not.

1360 Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience


FOCUS | Review
| Article
NATure NeurosCIenCe FOCUShttps://doi.org/10.1038/s41593-018-0242-x
Review Article
a
Nurturer:
TGFβr (homeostasis)
Chemokine receptors (migration)
C1q (synaptic remodeling)
Trem2 (apoptotic neuron removal)

Sentinel:
Sensome

c
Warrior:
Phagocytic receptors
Antimicrobial peptides
Pattern-recognition receptors
RNS, ROS production

Fig. 2 | Three proposed functional states of microglia. a, Nurturer state: microglia (left) stained for Cd11b (brown) in a normal brain are highly ramified
and evenly spaced throughout the brain parenchyma. In their nurturer role they maintain milieu homeostasis, participate in synaptic remodeling and
migration, and remove apoptotic neurons, all mediated by specific receptors and receptor-linked pathways. b, Sentinel state: micrograph taken from a
video using two-photon microscopy from a Cx3cr1-GFP mouse with a cranial window shows a cluster of green microglia with abundant processes. The
video from which this micrograph was taken (Supplementary Video 1) shows that microglia (green) processes are in constant motion, surveilling their
surroundings. Focal laser-induced injury initiates microglia response, with those microglia closest to the site of injury displaying polarization of surveilling
processes toward the area of injury. Microglia sensing is mediated by proteins encoded by sensome genes, which are portals for microglia to perform
their housekeeping and host-defense functions. c, Warrior state: microglia (left) stained for Cd11b (brown) accumulate around Aβ​deposits stained with
thioflavin-S (green), where they are observed to be two- to fivefold denser than in neighboring areas. The warrior morphology becomes stockier and
less ramified, and defense against infectious pathogens and injurious-self proteins including Aβ​is mediated through microglial Fc receptors, TLRs, viral
receptors, and antimicrobial peptides. Sensing is a prerequisite for microglia to perform their housekeeping and host-defense functions.

Aβ clearance. Aβ​deposition is regulated by equilibrium between speck-like protein containing a CARD (ASC) which binds Aβ​, caus-
Aβ​production and clearance. Small changes in this equilibrium ing its aggregation and leading to further amyloid ‘seeding’ and
result in abnormal accumulation. Aβ​clearance involves, in part36, spreading of amyloid pathology42. Similarly, Aβ​-induced cytokines
phagocytosis and endocytosis via microglial scavenger recep- promote tau hyperphosphorylation and pathology, thus initiating a
tors (SRs)37,38 and extracellular degradation by Aβ​-degrading self-perpetuating loop that culminates in worsening disease44,45. The
enzymes6,36. Decreased clearance contributes to Aβ​accumulation in double-edged sword metaphor refers to various stages of a single
late-onset AD. In support of this concept, microglia from a mouse microglia in AD. At a certain timepoint during disease progres-
model of Aβ​deposition (Aβ​-mice) have reduced expression of sion, microglia assume a useful role, then progress into a dysfunc-
Aβ​-phagocytic receptors and Aβ​-degrading enzymes, but their abil- tional cell which ultimately becomes deleterious. In support of this
ity to produce proinflammatory cytokines was maintained6. These concept, recent transcriptomic studies of microglia in normal and
results suggest that Aβ​accumulation is in part due to failure of Aβ​-mice identified subpopulations defined as disease-associated
microglia to clear this toxic peptide. microglia (DAM)8,24. DAMs are located around Aβ​plaques and
have dysregulated expression of sensing, housekeeping, and host-
Aβ-induced inflammation. Microglia–Aβ​interactions lead to defense genes. It is not clear how DAMs differ from ‘dark microglia’
early synapse loss39, production of neurotoxic reactive oxygen and associated with Aβ​deposits, which exhibit condensed cytoplasms
nitrogen species (ROS and RNS), NLRP3 inflammasome activa- and nucleoplasms and express high levels of CD11b and Trem246.
tion, and production of proinflammatory cytokines and TNF27,40–42. These findings support a direct link between aberrant microglial
This requires Aβ​interaction with microglial pattern recognition functions and AD and suggest that a subset of microglia transition
receptors (PRRs) including TLRs, SRs, and complement receptor from a homeostatic to DAMs in AD.
3 (CR3)7,43.
Tauopathies
Microglia in AD, a double-edged sword. Based on these findings, Tauopathies are neurodegenerative diseases characterized by hyper-
microglial–Aβ​interaction is a double-edged sword. While monitor- phosphorylated and aggregated tau and neurofibrillary tangles.
ing the brain environment, microglial sensing of Aβ​peptides results Tauopathies include AD, progressive supranuclear palsy, cortico-
in Aβ​clearance and removal of the injurious agent (Fig. 3). However, basal degeneration, frontotemporal dementia (FTD), and chronic
persistent production of Aβ​and its chronic interaction with traumatic encephalopathy (CTE) associated with repetitive con-
microglia drive further amyloid deposition. Indeed, Aβ​-induced cussive head injuries. Most tauopathies have neuronal and glial
proinflammatory cytokines reduce microglial Aβ​clearance abil- accumulations of toxic insoluble tau, neuronal loss, and proinflam-
ity, and NLRP3 activation releases microglial apoptosis-associated matory microglia47.

Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience 1361


Review Article | FOCUS
Review Article | FOCUS
https://doi.org/10.1038/s41593-018-0242-x NATure NeurosCIenCe
Non-cell-autonomous Microglia-autonomous

C9orf72
Disease stimulus Sensing Host defense TDP-43
Progranulin
Trem2
AD Aβ, tau
SRs, MMPs, AβDE Phagocytosis
ALS mSOD SRs
Degradation
FTD Tau
FTLD Tau Neuroinflammation
Inflammasome
CTE Tau
Activation of neuronal killing pathways
CJD PrPsc
mSOD
PD α-Synuclein C9orf72
PRRs, ROS, RNS
HD HTT HTT

Microglia-autonomous

Fig. 3 | Effectors of microglia function associated with neurodegeneration. Two common themes for microglia’s roles in neurodegenerative diseases
emerge. As microglia perform their normal sentinel function, they encounter aberrant or misfolded proteins such as Aβ​, aggregated α​-synuclein, oxidized
or mSOD1, or PrPsc. In response to these toxic stimuli, microglia perform their host-defense function, attempting to clear these agents via SRs and
other PRRs. The nature of the aberrant proteins or their persistent production disrupts microglial housekeeping functions and dysregulates microglial
host-defense functions, leading to an exaggerated proinflammatory response, neurotoxicity, and neurodegeneration. A second theme is that in some
neurodegenerative diseases, such as AD, ALS, and HD, mutations in specific genes cause self-autonomous dysregulation of host defense, thereby initiating
or exaggerating proinflammatory responses, resulting in neurotoxicity and neurodegeneration. For example, mutations in TDP-43, progranulin, and Trem2
affect phagocytosis and associated degradation pathways (purple), whereas mutations in mSOD and HTT affect inflammasome activation and neuronal
killing pathways (red). Mutations in C9orf72 affect both phagocytosis and inflammasome pathways. MMPs, matrix metalloproteases;
Aβ​DE, Aβ​-degrading enzymes.

Microglia can engulf, degrade, and clear tau48,49. In contrast, pathogenic pathways, raising the possibility that microglia may also
when activated, proinflammatory microglia increase tau phos- be a double-edged sword in PD.
phorylation50 and drive the spread of tau pathology48. This is
supported by human studies showing elevated levels of microves- Multiple sclerosis
icle-associated tau in the cerebrospinal fluid and blood of AD Multiple sclerosis (MS) patients have demyelinated plaques in the
patients and by brain imaging studies showing proinflammatory white and gray matter. Ongoing disease leads to progressive neu-
microglia in FTD, progressive supranuclear palsy, and CTE51–53. rodegeneration, resulting in brain atrophy. Neuroinflammation
Proinflammatory microglia precede visible tau pathology in is present in all stages of MS, and a proposed classification of MS
transgenic mice, and their activation is attenuated by the immu- lesions relies in part on the presence or absence of microglia in
nosuppressant drug FK506, which attenuates tau pathology and the lesions66.
extends lifespan in these mice, suggesting that microglia can Microglia in MS may be detrimental or beneficial. In experi-
mediate tau neurotoxicity54. mental autoimmune encephalomyelitis (EAE), a mouse model of
The double-edged sword concept of microglia’s role in AD also MS, microglia release proteases, proinflammatory cytokines, ROS,
applies to tauopathies (Fig. 3). Microglia first sense and clear tau in and RNS, and they recruit reactive T lymphocytes, thereby causing
an attempt to protect from tau toxicity, but dysregulation of microg- toxicity to neurons and oligodendrocyte precursors. Targeted dele-
lial sensing and housekeeping pathways, such as the Cx3cr1 and tion of the transforming growth factor (TGF)-β​-activated kinase 1
Trem2 pathways, lead to dysregulation of the host-defense pathway, in microglia in EAE reduced CNS inflammation and axonal and
resulting in a neuroinflammatory response gone awry, causing neu- myelin damage by cell-autonomous inhibition of the NF-κ​B, JNK,
ronal damage and loss55–57. and ERK1/2 pathways11. These results suggest that microglia pro-
mote tissue injury in EAE. However, at disease onset, microglia pro-
Parkinson’s disease mote axonal regeneration, remyelination, clearance of inhibitory
Parkinson’s disease (PD) is the second most common neurodegen- myelin debris, and the release of neurotrophic factors, suggesting
erative disease, affecting 1.2% of individuals over age 65. Most cases a beneficial role67.
are sporadic and 5–10% are inherited58. Neurodegeneration occurs These observations are likely to be relevant to MS stages, as
in the substantia nigra, with dopaminergic denervation of the microglia are closely associated with lesions with active demy-
striatum and accumulation of Lewy bodies containing aggregated elination68. It is possible that the detrimental and beneficial roles
α​-synuclein59. Reactive microglia expressing HLA-DR are abundant of microglia in MS depend on the stage of the disease or of spe-
in the substantia nigra of PD patients60. Positron-emission tomog- cific lesions. What is clear is that essential microglial functions are
raphy (PET) studies show widespread proinflammatory microglia, altered in EAE and possibly in MS, including their ability to sense
but this response does not correlate with clinical severity, suggesting and clear debris and mount a neuroprotective response.
it occurs early in the disease61.
The mechanism(s) by which microglia participate in PD may be Huntington’s disease
similar to those in AD (Fig. 3). Microglia internalize and degrade Huntington’s disease (HD) is characterized by progressive atrophy
α​-synuclein, possibly to clear it. A defect in this process leads to of the striatum and cortex in advanced disease69. Microscopically,
accumulation of extracellular α​-synuclein similar to Aβ​62. Microglia there are intranuclear inclusions containing the protein huntingtin
accumulate near α​-synuclein deposits and become proinflamma- (HTT) and neurodegeneration of medium-size spiny, enkephalin-
tory in a manner dependent on receptors that also bind Aβ​, such containing inhibitory neurons69. HD is caused by mutations in HTT
as CD36 and TLR263–65. These findings, which need to be validated (mHTT) that lead to expansion of the trinucleotide CAG stretch,
in animal models of PD, suggest that AD and PD have similar which translates into a polyglutamine stretch in the HTT protein69.

1362 Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience


FOCUS | Review
| Article
NATure NeurosCIenCe FOCUShttps://doi.org/10.1038/s41593-018-0242-x
Review Article
HTT mRNA is expressed in microglia at a relatively high level, host defense, may contribute to neurodegeneration in patients car-
similarly to TLR47. Proinflammatory microglia are seen early in HD rying the C9orf72 expansion (Fig. 3). Future functional studies with
by PET. In presymptomatic HD gene carriers, their presence cor- microglia from C9orf72 patients would help clarify their complex
relates with a higher probability of developing HD in 5 years70. In role in this subset of ALS.
HD patients, the presence of proinflammatory microglia correlates Transgenic mice expressing inducible human TDP-43 (hTDP-
with HD severity71,72. 43) exhibit progressive motor neuron loss but only subtle microglial
HD is associated with altered microglial function and mRNA changes88. Following suppression of hTDP-43 transgene expres-
profile. Expression of mHTT in microglia confers a cell-autonomous sion, microglia selectively cleared the existing neuronal hTDP-43.
increase in proinflammatory genes73 (Fig. 3). Levels and transcrip- When microgliosis was blocked during the early recovery phase
tional activities of the myeloid lineage-determining factors PU.1 using a CSF1R and c-Kit inhibitor, these mice failed to regain full
and CCAAT/enhancer-binding protein (C/EBP)-α​,β​are increased motor function, suggesting a neuroprotective role for microglia88.
and exhibit enhanced binding at thousands of genomic locations Interestingly, conditional deletion of TDP43 in microglia promotes
in HD patients and mouse models73. This correlates with increased their phagocytic functions and leads to enhanced synapse loss89.
expression of IL6 and TNF. These changes are unique to microg- While additional work is required to establish a clear pathway link-
lia and not observed in other myeloid cells73. Functionally, several ing TDP-43, microglia, and ALS pathogenesis, these findings sug-
of the genes that are increased in mHTT microglia are involved in gest dysregulation of microglial phagocytic function in ALS patients
sensing their milieu, such as Tlr2, Cd14, Fcgr1, Clec4d, Adora3, Tlr9, with TDP-43 mutations.
Tnfrsf1b, and others7,73, suggesting a possible increase in capacity to We propose that several genes with ALS-associated mutations
sense extracellular stimulants. This was associated with increased regulate microglial host defense functions, including production
microglial neurotoxicity73. Upregulation of IL6 and TNF mRNA73 of ROS (mutant SOD1), cytokines (C9orf72), and phagocytosis
suggest that the microglial response is an exaggerated host-defense (C9orf72 and TDP43; Fig. 3). These findings support a key role for
response to rid the brain of mHTT that went awry, thereby exacer- microglia in ALS pathogenesis but indicate that targeting microg-
bating neurodegeneration. lia for potential ALS therapy should be tailored to the specific
pathway(s) affected and that a gunshot approach is not a useful
Amyotrophic lateral sclerosis therapeutic strategy.
Most patients with amyotrophic lateral sclerosis (ALS; also called
Lou Gehrig’s disease) have sporadic ALS but ~10% of patients have Prion diseases
mutations in specific genes including SOD1, C9orf72, TDP43, and Prion diseases are genetic, sporadic, or acquired neurodegenerative
FUS74. The disease presents loss of motor neurons in the cortex, disorders resulting from sustained aggregation of PrPsc, the protein-
brainstem, and spinal cord. Microglia expressing proinflammatory ase-resistant form of the prion protein. Examples of prion diseases
markers are found near injured neurons in autopsies75 and seen by include Creutzfeldt–Jakob disease, scrapie, and chronic wasting
PET in the brains of live ALS patients76. disease. Prion-related neurodegeneration includes neuronal loss,
Transgenic mice overexpressing mutant human SOD1 (mSOD) increased proinflammatory microglia, and spongiform changes90.
develop a progressive motor neuron disease similar to ALS77. Microglia phagocytose PrPsc as early as 60 days postinfection91, and
mSOD1 expression in microglia accelerates disease onset78, and their depletion increases prion titers and susceptibility to prion
microglial activation exacerbates motor neuron death79. Microglia infection92, suggesting they help control prion disease.
change their phenotype with disease progression. Some proin- The double-edged sword metaphor also applies to microglia in
flammatory microglia are seen in the spinal cord before clinical prion diseases. Microglia produce ROS in response to the PrP106-126
disease develops, increase with disease progression, and persist fragment and enhance its neurotoxicity. Ablation of superoxide-
into end-stage disease80. Microglia isolated from mSOD1 mice at producing enzymes protected mice from PrP toxicity, further
disease-onset were neuroprotective, in contrast to microglia iso- suggesting that microglia mediate prion neurodegeneration93.
lated at end-stage disease81. Neurotoxicity of mSOD1 microglia is Upregulation of proinflammatory IL-1β​, IL-6, inducible nitric oxide
NF-κ​B-dependent82 and partly mediated by IL-1β​83. These findings synthase (iNOS), NF-κ​B, cyclophilin A, matrix metalloproteinases,
directly implicate microglia with mSOD1 in ALS progression. and NLRP3 inflammasome components have all been demon-
Pathways leading to microglial activation and neurotoxicity in strated in prion disease microglia94,95. Whether this proinflamma-
ALS are both cell-autonomous and dependent on exogenous stim- tory response affects disease progression is unclear, since deletion
uli (Fig. 3). Expression of mSOD1 in microglia disrupts regulation of NLRP3 or the inflammasome adaptor Pycard did not markedly
of NADPH oxidase, leading to excessive neurotoxic superoxide affect the clinical course of scrapie in mice94.
production84. Intraneuronal or extracellular misfolded SOD1 is Prion infection affects microglial sensing and housekeeping abil-
sensed by microglia similarly to their sensing of Aβ​ or α​-synuclein ity, in part by disrupting the Cx3cr1–fractalkine pathway96. PrPsc
through TLRs and SRs, rendering them proinflammatory85. These also reduces microglial phagocytosis of aberrant proteins, includ-
findings suggest that the two major microglial functions altered in ing PrPsc and apoptotic debris or cells, despite production of proin-
ALS include the sensing of exogenous stimuli and danger signals flammatory mediators, suggesting dysregulated host defense97. As
and the host response. They also point to shared neurodegenerative in AD, PD, and ALS, the effects of PrPsc on microglia appear to be
pathways between ALS, AD, and PD, as ALS microglia also change mediated by SRs and TLRs in an Src-kinase-dependent manner98,99.
their phenotype with disease progression from neuroprotective It is plausible that microglia try initially to clear PrPsc, leading to
to neurotoxic81. their persistent activation and dysregulated functions in another
Transgenic mice carrying the human C9orf72 gene with disease- example of a host-defense response gone awry, resulting in neuro-
associated expansion repeats display pathologic features of ALS, toxicity and subsequent disease progression.
without behavioral abnormalities or neurodegeneration86. In con- Two common themes for microglia’s roles in neurodegenerative
trast, mice deficient in C9orf72 exhibited enhanced TNF and IL-1 diseases emerge (Fig. 3). First, while microglia are performing their
production when stimulated and defective maturation of phago- normal sentinel function, they sense the presence of an aberrant or
somes to lysosomes87. While these findings seem contradictory, they misfolded protein such as Aβ​, aggregated α​-synuclein, oxidized or
suggest that C9orf72 is required for normal microglial function and mutant SOD1, or PrPsc. In response to these toxic stimuli, microg-
that altering microglial ability to clear aggregated proteins by alter- lia perform their host-defense function, attempting to clear these
ing phagosomes-to-lysosomes maturation, an important step in agents via SRs and other PRRs. The nature of the aberrant proteins

Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience 1363


Review Article | FOCUS
Review Article | FOCUS
https://doi.org/10.1038/s41593-018-0242-x NATure NeurosCIenCe

Indirect
pathways Release of TNF-α, or reduced production of
nutritive BDNF and IGF

Neuronal death
Direct
pathways
Release of matrix metalloproteases

NADPH NO RNS
ROS Functional alterations in cellular lipids, proteins, and DNA
Induce neurons to transiently flip their membrane
phosphatidylserine ‘eat me’ signal for microglia
Overexpression of iNOS Inhibiting mitochondrial cytochrome oxidase
Release of glutamate Excitotoxic neuronal death

Fig. 4 | How microglia damage or kill neurons: There are several direct and indirect tools used by microglia to perform this task. When microglia interact
with ligands such as an infectious pathogen, Aβ​, PrPSc, aggregated α​-synuclein, or mSOD1, several pathways are activated. NADPH oxidase produces
superoxide and derivative oxidants. Nitric oxide and its derivatives are produced by iNOS. Glutamate, cathepsin B, and other proteases are released, or
phagocytic killing of stressed neurons occurs. Oxidative lipid damage reduces membrane fluidity and membrane potential and increases ion permeability,
resulting in organelle swelling, loss of membrane depolarization, and rupture of the plasma membrane, leading to necrosis. Microglia also utilize indirect
means to kill and damage neurons, including release of TNF, which stimulates NMDA receptor activity, or reduced production of nutritive BDNF and IGF.

or their persistent production disrupts microglial housekeeping Trem2


functions and dysregulates microglial immune checkpoints and The Trem2 gene encodes an innate immune receptor of the immu-
pathways that keep inflammation in check, such as the Cx3cr1 or noglobulin family located on chromosome 6 in humans and chro-
progranulin pathways, leading to an exaggerated proinflammatory mosome 17 in mouse105. Trem2 is expressed on macrophages,
response, neurotoxicity, and neurodegeneration. A second theme dendritic cells, osteoclasts, and microglia and is part of the microg-
is that in some neurodegenerative diseases, mutations in specific lial sensome7. Trem2 ligands include ApoE, phosphatidylserine,
genes, such as Trem2, HTT, and TDP43, cause a self-autonomous sphingomyelin, Aβ​, dead neurons, and damaged myelin106–108.
dysregulation of sensing, housekeeping, or host defense, thereby Trem2 forms a signaling complex with the adaptor tyrosine
initiating or exaggerating proinflammatory responses and leading kinase-binding protein (TyroBP or DAP12)105. Ligand binding to
to neurotoxicity and neurodegeneration. Trem2 triggers phagocytosis and chemotaxis and negatively regu-
lates TLR-induced inflammatory responses105. The extracellular
Common pathways to neurodegeneration: microglial domain of Trem2 can be released as a soluble protein (sTrem2) and
immune checkpoints increases with age and with MS, AD, and FTD109. How is Trem2
How do microglia damage and kill neurons?. A recurring theme involved in specific neurodegenerative disease?
in neurodegenerative diseases is the damaging and killing of neu-
rons by microglia using several direct and indirect tools (Fig. 4). AD. Genome-wide association studies identified mutations in Trem2
When activated by ligands such as an infectious pathogen, Aβ​, PrPSc, as major risk factors in late-onset AD35. Trem2 expression increases in
aggregated α​-synuclein, or mSOD1, NADPH produces superoxide, plaque-associated microglia and infiltrating monocytes, suggesting
which is released and either transformed into H2O2 by extracellular a role in the microglial response to Aβ​105. Trem2 deletion decreased
SOD or reacts with NO to produce peroxynitrite100. These cause cel- microglial phagocytosis, proliferation, and survival, and increased
lular necrosis100 or apoptosis101 (Fig. 4). Microglia also cause exci- proinflammatory cytokines and RNS110. Trem2−/−Aβ​-mice have
totoxic neuronal death either by overexpressing iNOS or directly increased Aβ​deposition, reduced numbers of myeloid cells around
releasing glutamate101,102. Microglial proteases such as cathepsins plaques, reduced Aβ​phagocytosis, and greater neuritic dystrophy
are released in response to Aβ​, leading to neuronal apoptosis103, in early disease111. This reduction of microglia around plaques was
and matrix metalloproteases can cause neuronal injury in hypoxia- attributed to lower proliferation, decreased metabolic fitness112, and
ischemia (Fig. 4)104. Microglia can also damage neurons indirectly, increased death111. Since microglia form a physical barrier to prevent
either by releasing TNF or by reducing production of nutritive or plaque expansion and protect neurons113, increased accumulation of
neuroprotective brain-derived neurotrophic factor (BDNF) and dystrophic neurons in Trem2−/−Aβ​-mice was attributed to decreased
insulin-like growth factor (IGF), thereby increasing neuronal apop- clearance by microglia rather than increased neuronal death.
tosis (Fig. 4)101. Collectively, these results implicate Trem2 in microglial recruitment
It is evident that the microglial host-defense function provides to Aβ​plaques and restricting exposure of neurons to toxic Aβ​.
microglia with the tools for fratricide to become neuronal killers. Trem2 variants comprise amino acid substitutions, frameshift
This does not happen continuously because microglia have several and nonsense mutations, and splice site alterations that likely result
immunological checkpoints or pathways that prevent their overreac- in loss of function114. The Trem2 variant with the strongest AD
tion to external stimuli. These include the Trem2, Cx3cr1–fractalkine association is R47H (three to four times increased risk), a single
sensing, and housekeeping pathways, the progranulin pathways that amino acid substitution in the extracellular domain114. In transgenic
keep their inflammatory response in check, and the SR pathways Aβ​-mice in which endogenous Trem2 was replaced with the human
that promote clearance of injurious stimuli. Dysregulation of any common variant of normal Trem2 or the variant R47H115, RH47 was
of these pathways or disruption of the sentinel and housekeeping associated with reduced Aβ​-induced microglial responses, further
functions initiates or exacerbates neurodegeneration (Box 1). supporting that R47H reduces Trem2 function in vivo.

1364 Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience


FOCUS | Review
| Article
NATure NeurosCIenCe FOCUShttps://doi.org/10.1038/s41593-018-0242-x
Review Article
Box 1 | Microglia immune checkpoints disrupted in deficits compared with controls. Additional studies are needed
neurodegeneration to definitively assess the role of Trem2 in tauopathies and recon-
cile conflicting results.
As with lymphocytes, microglia have several immunological The association of Trem2 variants with AD, NHD, FTD, and
checkpoints that prevent their overreaction to external stimuli. possibly ALS and PD suggest that Trem2 is a major immune
These checkpoints are different from those of lymphocytes. They checkpoint that regulates an important microglial homeostatic
include the microglial Trem2, Cx3cr1–fractalkine sensing, and pathway. Disruption of this pathway by Trem2 variants leads to
housekeeping pathways and the progranulin pathways, which neurodegeneration (Box 1). However, the diversity of pathologies
keep their inflammatory response in check. Dysregulation of any associated with these diseases suggests that additional genetic
of these pathways or disruption of the sentinel, defense, or house- and/or epigenetic factors interact with Trem2 to cause a specific
keeping functions initiates or exacerbates neurodegeneration. disorder. Therefore, understanding the biology of Trem2 and how
Trem2 regulates all three microglial functions (sensing, it regulates microglial functions will be useful not only for help-
housekeeping, and host defense) and dysregulation of Trem2 ing patients with Trem2 variants but for understanding the role of
increases the risk for AD, FTLD, FTD, and possibly ALS. microglia in neurodegeneration in general. Ongoing work should
Cx3cr1 regulates sensing and housekeeping functions. determine whether Trem2 is a realistic therapeutic target for neu-
However, disruption of this pathway has not been shown to rodegenerative diseases, at what stage would therapy work best,
increase the risk for neurodegenerative diseases, but it alters and whether sTrem2 is a potential biomarker for efficacy.
disease courses in animal models.
The progranulin pathway regulates housekeeping and its Cx3cr1–fractalkine
disruption may increase risk for AD, FTLD, FTD, and ALS. Interactions between neuronal fractalkine (Cx3cl1) and its microg-
As with T cell immune checkpoints, microglial immune lial receptor, Cx3cr1, define another immune checkpoint and
checkpoints may be important targets for therapy. pathway that regulates microglial function, likely because fractal-
kine and Cx3cr1 promote reciprocal neuron–microglial signaling.
Membrane-bound fractalkine or its soluble form, (s)fractalkine, are
Single-cell RNA-seq identified three microglial subpopulations the only known ligands for Cx3cr1124.
in AD: homeostatic, intermediate, and DAM8. Transition from Fractalkine–Cx3cr1 interactions regulate microglial homeostatic
homeostatic to the DAM state appears to be a two-step process functions and temper microglial response to inflammatory and
characterized by a Trem2-independent homeostatic-to-interme- injurious stimuli. Blocking such interactions upregulates microg-
diate state and a Trem2-dependent intermediate-to-DAM state. lial TNF production, causing neurotoxicity125. In neurons, these
DAMs are neuroprotective and clear Aβ​. Another subtype of DAM interactions regulate synaptic maturation22 and promote neuronal
identified in aged mice, models of AD, ALS, and MS, is dependent survival125. However, as Cx3cr1 is also expressed on perivascular
on a Trem2–ApoE pathway and is induced by phagocytosis of apop- macrophages, studies using Cx3cr1−/− mice should be cautiously
totic neurons24. This subtype exhibits decreased housekeeping and interpreted; the effects observed may also be attributed to perivas-
sensing genes and increased neurodegeneration-associated genes, cular macrophages.
suggesting that the Trem2–ApoE pathway regulates a switch from
the homeostatic to a neurodegenerative phenotype. More work AD. In Aβ​and tau mouse models, Cx3cr1 deficiency enhances
is needed to reconcile the role of Trem2 in each DAM subtype. Aβ​ clearance126 with a gene-dosage effect, but worsens tau
However, there is consensus that Trem2 is a key regulator of microg- pathology55. In support of such dual role, overexpression of (s)
lia functions and phenotype in AD. fractalkine in tau mice but not in Aβ​-mice led to substantial
improvement127. The reason for this dichotomy is unclear, but
Nasu–Hakola disease. Trem2 variants are also associated with since Cx3cr1 is important for microglial sensing and homeostatic
increased risk for polycystic lipomembraneous osteodysplasia with functions, dysregulation of these functions is implicated. Because
sclerosing leukoencephalopathy, also called Nasu–Hakola disease of this dichotomy, the role of fractalkine–Cx3cr1 in AD remains
(NHD)116. NHD patients develop axonal degeneration, loss of white ambiguous and needs to be tested in mice with combined Aβ​ and
matter, and cortical atrophy. They exhibit increased microglial tau pathology.
density and activation suggesting a dysregulated proinflammatory
response117. PD. In the MPTP (a dopaminergic neurotoxic compound) model
of PD, Cx3cr1−/− mice showed more extensive neuronal loss than
Tauopathies, ALS, and PD. Trem2 may regulate tau pathology. Cx3cr1+/− mice128. In support of this neuroprotective role, over-
sTrem2 correlates with tau levels in cerebrospinal fluid early in expression of (s)fractalkine reversed MPTP toxicity for dopami-
clinical AD118 and with tangle score and paired helical filament lev- nergic neurons127. In contrast, in transgenic mice overexpressing
els in postmortem AD brains119. Trem2 variants were identified in α​-synuclein, Cx3cr1−/− microglia were less activated by aggregated
families with FTD or frontotemporal lobar degeneration (FTLD; α​-synuclein, leading to reduced loss of dopaminergic neurons129.
Supplementary Table 1). Studies assessing the risk of FTD or FTLD In mice deficient in LRRK2, another PD-associated gene130, Cx3cr1
in nonfamilial cases are conflicting, though Trem2 variants appear was upregulated and microglia had reduced responsiveness to lipo-
to influence clinical manifestations of these diseases120. One study polysaccharide. These studies suggest that altered Cx3cr1-mediated
found R47H to be a significant risk for ALS121, but this needs to be microglial sensing and housekeeping functions play a direct
replicated. Attempts to correlate R47H with increased PD risk have effect in PD.
had mixed results120.
Studies in mice are also conflicting. In one study, Trem2 ALS. In the mSOD1G93A mouse model of ALS, Cx3cr1−/− mice
deficiency exacerbated tau pathology, neurodegeneration, and showed more neuronal loss than Cx3cr1+/− or Cx3cr1+/+ mice, indi-
spatial learning deficits122. In another study, Trem2 deficiency cating a gene-dosage neuroprotective effect128. In support of this
reduced microgliosis, brain atrophy, and levels of inflamma- role, the loss of function of Cx3cr1 seen with the V249I rs3732379
tory cytokines without affecting tau levels123. Similarly, in a CTE Cx3cr1 variant was associated with accelerated progression and
model, injured Trem2−/− mice showed less hippocampal atrophy, reduced survival in some ALS patients131, but not in others132. In
reduced inflammatory transcripts, and improved behavioral both studies, Cx3cr1 variants did not increase risk for ALS. Though

Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience 1365


Review Article | FOCUS
Review Article | FOCUS
https://doi.org/10.1038/s41593-018-0242-x NATure NeurosCIenCe

these opposing findings are confusing, they suggest that Cx3cr1 is a Aβ​, mSOD1, aggregated α​-synuclein, and PrPsc, possibly to clear
disease-modifying factor in ALS and raise the need for studies with them. With disease progression, SR expression decreases, leading to
more patients. defective clearing and subsequent accumulation of these misfolded
proteins. Some of these receptors, such as CD36, form receptor
Prion diseases. Scrapie-infected hamsters’ brains show progres- complexes with other PRRS such as TLRs and initiate an inflam-
sive downregulation of fractalkine96. The same effect is observed matory response that leads to neurotoxicity and neurodegeneration.
in response to PrP106-126 in mixed neuron–glia cultures96. Infection
of Cx3cr1−/− mice reduced their time to disease-onset compared to Progranulin
wild-type mice, suggesting that Cx3cr1 may be protective in prion Progranulin is a secreted glycoprotein with neuro-immunomodu-
disease. No differences were seen in the pattern and localization of latory properties and autocrine neurotrophic functions important
microglia or in chemokine and cytokine levels133. This was not fully for long-term neuronal survival141. Progranulin deficiency leads to
duplicated in Cx3cr1−/− mice with a different genetic background134. age-dependent, progressive upregulation of lysosomal and innate
Repeating these studies and using fractalkine-deficient (Cx3cl1−/−) immunity genes, increased complement production, and enhanced
mice would help clarify the discrepancy. C1q-dependent synaptic pruning by microglia, suggesting that pro-
There is consensus that the Cx3cr1–fractalkine immune check- granulin is an important immune checkpoint that suppresses aber-
point is overall a neuroprotective pathway that regulates sensing, rant microglial activation in aging21. Dysregulation of this pathway
housekeeping, and host-defense functions of microglia (Box 1). in microglia may occur in neurodegenerative diseases. Interestingly,
However, there are multiple knowledge gaps in our understanding similar results were obtained in a mouse model of adrenomyeloneu-
of the roles of this pathway in neurodegeneration. ropathy. Deficiency in the very-long-chain fatty-acid transporter
ABCD1 increased complement activation and synapse loss and
Scavenger receptors aggravated microglial phagocytosis of neurons, suggesting a novel
SRs are innate immune PRRs that promote removal of non-self or pathway for ‘death by microglia’142.
altered-self ligands and elimination of degraded or harmful sub-
stances135. They have roles in AD, PD, ALS, and prion diseases. FTLD and AD. Haploinsufficiency caused by autosomal dominant
mutations in the progranulin gene leads to FTLD141. Progranulin
AD. SR-A1 is an Aβ​-phagocytic receptor37,38 expressed on microglia polymorphism is also linked to late-onset AD143. In Aβ​-mice, pro-
surrounding Aβ​plaques in humans and Aβ​-mice. SR-A1-deficiency granulin overexpression inhibits Aβ​deposition and protects from
decreases microglial Aβ​uptake by 60%37,136 and increases mortal- Aβ​toxicity. Selective reduction of microglial expression of progran-
ity and Aβ​ accumulation37,136, whereas upregulation of SR-A leads ulin in Aβ​-mice also impairs phagocytosis, increases plaque load,
to reduced Aβ​burden in Aβ​-mice37. Therapeutically, it would be and exacerbates cognitive deficits. Thus, increasing progranulin
beneficial to upregulate microglial expression of SRA1 to increase expression is proposed as a potential therapy for FTLD and AD144.
Aβ​ clearance.
SR-B2 (also called CD36) is also a microglial Aβ​receptor that PD. Progranulin polymorphism is associated with increased PD
forms a complex with TLR-4 and TLR-6, leading to cytokine, che- risk but the mechanism is not clear143. In the MPTP mouse model
mokine, and ROS production, as well as microglial migration, of PD, upregulation of progranulin expression in nigrostriatal neu-
inflammasome activation, and neurotoxicity27,40. Polymorphism rons was accompanied by reductions in markers of MPTP-induced
(rs3211892) that increases CD36 levels is associated with increased inflammation and apoptosis, and it protected these neurons from
AD risk137. Therapeutically, it may be advantageous to inhibit CD36– MPTP toxicity and preserved striatal dopamine content and turn-
Aβ​binding or signaling to reduce microglial neurotoxicity138. over. This was associated with preservation of locomotor function.
Binding of Aβ​to the microglial SR-J1 (also called RAGE) acti- Dysregulation of the progranulin pathway may therefore contrib-
vates MAPK and NF-κ​B and contributes to synaptic dysfunction ute to PD pathogenesis, and restoring progranulin levels may have
through IL-1β​release in 2- to 3-month-old animals139. Six-month- therapeutic potential for PD145.
old RAGE−/− Aβ​-mice have reduced Aβ​deposition and more Aβ​
-degrading enzymes140. However, no change in cognition or microg- ALS. Five progranulin mutations (four missense and one 5′​ regu-
lial recruitment to plaques was seen in 12-month-old mice, suggest- latory variant) are associated with reduction in age at onset and
ing that RAGE is not essential for microglial recruitment, but could shorter survival after onset of ALS, suggesting that progranulin
affect Aβ​processing in early disease140. modifies ALS disease course146. The mechanism of this associa-
These reports indicate that various SRs play complementary tion is not clear. Based on mouse and human studies, progranulin
roles in microglia–Aβ​interactions that may have therapeutic impli- appears to be an important immune checkpoint that modulates the
cations for AD. Pharmacologic upregulation of SR-A1 expression microglial response to external stimuli and prevents them from
or function may be helpful for AD treatment, whereas blocking overreacting to such stimuli and thereby causing neurodegenera-
CD36– or RAGE–Aβ​interactions or reducing expression of these tion (Box 1).
SRs may stop or delay AD progression.
Peripheral regulation of microglia in neurodegeneration
PD, ALS, and prion diseases. Similarly to the roles they play in AD, Despite the anatomical separation between the CNS and the gas-
SRs may also be involved in PD, ALS, and prion diseases through trointestinal system, a bidirectional connection between the two,
their interactions with misfolded proteins. SR-B2 promotes α​-synu- known as the ‘brain–gut axis’ exists. Intestinal bacterial coloniza-
clein–microglial interactions64 and senses misfolded mutant or oxi- tion regulates immune system maturation and development in the
dized SOD1 in diseased motor neurons or in the extracellular space, CNS. Eradication of gut microbiota also alters microglial num-
rendering these cells proinflammatory85. Aggregated prion peptides bers, size, transcriptomes, and surveillance functions and down-
also interact with microglial SRs, leading to uptake and/or neuro- regulates host-defense genes147. Recolonization with a complex
toxin production99. microbiota or with microbiota-derived products partially restores
Cumulatively, these reports suggest a potential common path- microglial properties147.
way for handling misfolded proteins in neurodegenerative diseases, Human studies show associations between gut microbiota and
involving SRs (Fig. 3). Because of their ability to bind a diverse class neurodegenerative diseases148. In Aβ​-mice, manipulating gut micro-
of ligands, SRs are important sensors of misfolded proteins including biota influences Aβ​deposition and alters microglial reactivity and

1366 Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience


FOCUS | Review
| Article
NATure NeurosCIenCe FOCUShttps://doi.org/10.1038/s41593-018-0242-x
Review Article
morphology by unknown mechanisms149. In a PD model, antibiotic 12. Kierdorf, K. et al. Microglia emerge from erythromyeloid precursors via
treatment ameliorates PD pathology, whereas oral administration Pu.1- and Irf8-dependent pathways. Nat. Neurosci. 16, 273–280 (2013).
13. Tay, T. L. et al. A new fate mapping system reveals context-dependent
of specific microbial metabolites promotes neuroinflammation and random or clonal expansion of microglia. Nat. Neurosci. 20, 793–803 (2017).
motor symptoms150. These studies add another layer to how microg- 14. Wang, Y. et al. IL-34 is a tissue-restricted ligand of CSF1R required for
lial functions are regulated in normal and diseased states, suggest- the development of Langerhans cells and microglia. Nat. Immunol. 13,
ing that gut-to-brain signaling modulates neurodegeneration. 753–760 (2012).
15. Abud, E. M. et al. iPSC-derived human microglia-like cells to study
neurological diseases. Neuron 94, 278–293.e9 (2017).
Conclusions 16. Muffat, J. et al. Efficient derivation of microglia-like cells from human
Understanding of microglial biology has increased exponentially pluripotent stem cells. Nat. Med. 22, 1358–1367 (2016).
in the past decade. Microglial gene expression profiles are being 17. Ryan, K. J. et al. A human microglia-like cellular model for assessing the
defined at the single-cell level and correlated with specific func- effects of neurodegenerative disease gene variants. Sci. Transl. Med. 9,
tions. We are also beginning to understand microglial roles in neu- eaai7635 (2017).
18. Gosselin, D. et al. An environment-dependent transcriptional network
rodegeneration and to explore pathways that regulate their response specifies human microglia identity. Science 356, eaal3222 (2017).
to injury, including neuroprotective immune checkpoint pathways 19. Fourgeaud, L. et al. TAM receptors regulate multiple features of microglial
that keep the microglial proinflammatory response in check and physiology. Nature 532, 240–244 (2016).
pathways that promote clearance of injurious stimuli. We are also 20. Vasek, M. J. et al. A complement-microglial axis drives synapse loss during
exploring how peripheral influences from the gut microbiome can virus-induced memory impairment. Nature 534, 538–543 (2016).
21. Lui, H. et al. Progranulin deficiency promotes circuit-specific synaptic
alter progression of such injury. Yet significant knowledge-gaps exist. pruning by microglia via complement activation. Cell 165, 921–935 (2016).
Analyzing the transcriptomes and epigenetic profiles in various dis- 22. Zhan, Y. et al. Deficient neuron-microglia signaling results in impaired
ease states, understanding how aging and disease progression alter functional brain connectivity and social behavior. Nat. Neurosci. 17,
these profiles at the single-cell level, and correlating such changes 400–406 (2014).
with microglial behavior are necessary. Expanding the studies from 23. Fuhrmann, M. et al. Microglial Cx3cr1 knockout prevents neuron loss in
a mouse model of Alzheimer’s disease. Nat. Neurosci. 13, 411–413 (2010).
mouse models to human patients remains a major limiting factor 24. Krasemann, S. et al. The TREM2-APOE pathway drives the transcriptional
and requires development of new reliable in vitro cellular models phenotype of dysfunctional microglia in neurodegenerative diseases.
derived from patient samples as well as additional technologies for Immunity 47, 566–581.e9 (2017).
in vivo imaging and analysis. In this regard, the ability to isolate 25. Healy, L. M. et al. MerTK is a functional regulator of myelin phagocytosis
primary microglia from fresh postmortem brain tissues or to repro- by human myeloid cells. J. Immunol. 196, 3375–3384 (2016).
26. Liddelow, S. A. et al. Neurotoxic reactive astrocytes are induced by activated
gram human stem cells or monocytes to develop into microglia-like microglia. Nature 541, 481–487 (2017).
cells, as well as the ability to generate such cells from patients with 27. El Khoury, J. B. et al. CD36 mediates the innate host response to
various neurodegenerative diseases, are steps in the right direc- beta-amyloid. J. Exp. Med. 197, 1657–1666 (2003).
tion15–17. However, since correct programming is dependent on the 28. El Khoury, J. et al. Ccr2 deficiency impairs microglial accumulation
cell’s environment18, new techniques to incorporate these cells into and accelerates progression of Alzheimer-like disease. Nat. Med. 13,
432–438 (2007).
three-dimensional organoids or improving existing techniques for 29. Selkoe, D. J. & Hardy, J. The amyloid hypothesis of Alzheimer’s disease at
organotypic brain slice culture while maintaining their in vivo tran- 25 years. EMBO Mol. Med. 8, 595–608 (2016).
scriptomic and epigenetic profiles are crucial next breakthroughs 30. Frautschy, S. A. et al. Microglial response to amyloid plaques in APPsw
waiting to be achieved. These steps are necessary for any successful transgenic mice. Am. J. Pathol. 152, 307–317 (1998).
effort to harness the power of microglia for treatment of neurode- 31. D’Andrea, M. R., Cole, G. M. & Ard, M. D. The microglial phagocytic role
with specific plaque types in the Alzheimer disease brain. Neurobiol. Aging
generative disease. 25, 675–683 (2004).
32. Tooyama, I., Kimura, H., Akiyama, H. & McGeer, P. L. Reactive microglia
Received: 12 February 2018; Accepted: 6 August 2018; express class I and class II major histocompatibility complex antigens in
Published online: 26 September 2018 Alzheimer’s disease. Brain Res. 523, 273–280 (1990).
33. Martin, E., Boucher, C., Fontaine, B. & Delarasse, C. Distinct inflammatory
References phenotypes of microglia and monocyte-derived macrophages in Alzheimer’s
1. Río Hortega, P. Noticia de un nuevo y fácil método para la coloración disease models: effects of aging and amyloid pathology. Aging Cell 16,
de la neuroglia y el tejido conjuntivo. Trab. Lab. Invest. Biol. 15, 27–38 (2017).
367–378 (1918). 34. Lambert, J. C. et al. Meta-analysis of 74,046 individuals identifies 11 new
2. Giulian, D. & Baker, T. J. Characterization of ameboid microglia isolated susceptibility loci for Alzheimer’s disease. Nat. Genet. 45, 1452–1458 (2013).
from developing mammalian brain. J. Neurosci. 6, 2163–2178 (1986). 35. Jonsson, T. et al. Variant of TREM2 associated with the risk of Alzheimer’s
3. Jung, S. et al. Analysis of fractalkine receptor CX(3)CR1 function by disease. N. Engl. J. Med. 368, 107–116 (2013).
targeted deletion and green fluorescent protein reporter gene insertion. 36. El Khoury, J. & Hickman, S.E. Mechanisms of amyloid-beta clearance in
Mol. Cell. Biol. 20, 4106–4114 (2000). Alzheimer’s disease. in Research Progress in Alzheimer’s Disease and
4. Haynes, S. E. et al. The P2Y12 receptor regulates microglial activation by Dementia, vol. 4. (ed. Sun, M.-K.) 37–66 (Nova Science Publishers,
extracellular nucleotides. Nat. Neurosci. 9, 1512–1519 (2006). Hauppauge, NY, USA, 2009).
5. Yang, G., Parkhurst, C. N., Hayes, S. & Gan, W. B. Peripheral elevation of 37. Frenkel, D. et al. Scara1 deficiency impairs clearance of soluble amyloid-β​
TNF-α​leads to early synaptic abnormalities in the mouse somatosensory by mononuclear phagocytes and accelerates Alzheimer’s-like disease
cortex in experimental autoimmune encephalomyelitis. Proc. Natl. Acad. progression. Nat. Commun. 4, 2030 (2013).
Sci. USA 110, 10306–10311 (2013). 38. El Khoury, J. et al. Scavenger receptor-mediated adhesion of microglia to
6. Hickman, S. E., Allison, E. K. & El Khoury, J. Microglial dysfunction and beta-amyloid fibrils. Nature 382, 716–719 (1996).
defective beta-amyloid clearance pathways in aging Alzheimer’s disease 39. Hong, S. et al. Complement and microglia mediate early synapse loss in
mice. J. Neurosci. 28, 8354–8360 (2008). Alzheimer mouse models. Science 352, 712–716 (2016).
7. Hickman, S. E. et al. The microglial sensome revealed by direct RNA 40. Coraci, I. S. et al. CD36, a class B scavenger receptor, is expressed on
sequencing. Nat. Neurosci. 16, 1896–1905 (2013). microglia in Alzheimer’s disease brains and can mediate production of
8. Keren-Shaul, H. et al. A unique microglia type associated with restricting reactive oxygen species in response to beta-amyloid fibrils. Am. J. Pathol.
development of Alzheimer’s disease. Cell 169, 1276–1290.e17 (2017). 160, 101–112 (2002).
9. Lawson, L. J., Perry, V. H., Dri, P. & Gordon, S. Heterogeneity in the 41. Gold, M. & El Khoury, J. β​-amyloid, microglia, and the inflammasome in
distribution and morphology of microglia in the normal adult mouse brain. Alzheimer’s disease. Semin. Immunopathol. 37, 607–611 (2015).
Neuroscience 39, 151–170 (1990). 42. Venegas, C. et al. Microglia-derived ASC specks cross-seed amyloid-β​ in
10. El Khoury, J. Neurodegeneration and the neuroimmune system. Nat. Med. Alzheimer’s disease. Nature 552, 355–361 (2017).
16, 1369–1370 (2010). 43. Hickman, S. E. & El Khoury, J. The neuroimmune system in
11. Ransohoff, R. M. & El Khoury, J. Microglia in health and disease. Alzheimer’s disease: the glass is half full. J. Alzheimers Dis. 33(Suppl 1),
Cold Spring Harb. Perspect. Biol. 8, a020560 (2015). S295–S302 (2013).

Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience 1367


Review Article | FOCUS
Review Article | FOCUS
https://doi.org/10.1038/s41593-018-0242-x NATure NeurosCIenCe
44. Oddo, S., Caccamo, A., Kitazawa, M., Tseng, B. P. & LaFerla, F. M. Amyloid 77. Gurney, M. E. et al. Motor neuron degeneration in mice that express
deposition precedes tangle formation in a triple transgenic model of a human Cu,Zn superoxide dismutase mutation. Science 264,
Alzheimer’s disease. Neurobiol. Aging 24, 1063–1070 (2003). 1772–1775 (1994).
45. Villemagne, V. L. et al. Aβ​-amyloid and tau imaging in dementia. Semin. 78. Yamanaka, K. et al. Mutant SOD1 in cell types other than motor neurons
Nucl. Med. 47, 75–88 (2017). and oligodendrocytes accelerates onset of disease in ALS mice. Proc. Natl.
46. Bisht, K. et al. Dark microglia: a new phenotype predominantly associated Acad. Sci. USA 105, 7594–7599 (2008).
with pathological states. Glia 64, 826–839 (2016). 79. Apolloni, S., Amadio, S., Montilli, C., Volonté, C. & D’Ambrosi, N. Ablation
47. Ferrer, I. et al. Glial and neuronal tau pathology in tauopathies: of P2X7 receptor exacerbates gliosis and motoneuron death in the
characterization of disease-specific phenotypes and tau pathology SOD1-G93A mouse model of amyotrophic lateral sclerosis. Hum. Mol.
progression. J. Neuropathol. Exp. Neurol. 73, 81–97 (2014). Genet. 22, 4102–4116 (2013).
48. Asai, H. et al. Depletion of microglia and inhibition of exosome synthesis 80. Hall, E. D., Oostveen, J. A. & Gurney, M. E. Relationship of microglial and
halt tau propagation. Nat. Neurosci. 18, 1584–1593 (2015). astrocytic activation to disease onset and progression in a transgenic model
49. Bolós, M. et al. Direct evidence of internalization of tau by microglia in of familial ALS. Glia 23, 249–256 (1998).
vitro and in vivo. J. Alzheimers Dis. 50, 77–87 (2016). 81. Liao, B., Zhao, W., Beers, D. R., Henkel, J. S. & Appel, S. H. Transformation
50. Lee, D. C. et al. LPS- induced inflammation exacerbates phospho-tau from a neuroprotective to a neurotoxic microglial phenotype in a mouse
pathology in rTg4510 mice. J. Neuroinflammation 7, 56 (2010). model of ALS. Exp. Neurol. 237, 147–152 (2012).
51. Saman, S. et al. Exosome-associated tau is secreted in tauopathy models 82. Frakes, A. E. et al. Microglia induce motor neuron death via the
and is selectively phosphorylated in cerebrospinal fluid in early Alzheimer classical NF-κ​B pathway in amyotrophic lateral sclerosis. Neuron 81,
disease. J. Biol. Chem. 287, 3842–3849 (2012). 1009–1023 (2014).
52. Fiandaca, M. S. et al. Identification of preclinical Alzheimer’s disease by a 83. Meissner, F., Molawi, K. & Zychlinsky, A. Mutant superoxide dismutase
profile of pathogenic proteins in neurally derived blood exosomes: a 1-induced IL-1beta accelerates ALS pathogenesis. Proc. Natl. Acad. Sci. USA
case-control study. Alzheimers Dement. 11, 600–7.e1 (2015). 107, 13046–13050 (2010).
53. Cherry, J. D. et al. Microglial neuroinflammation contributes to tau 84. Harraz, M. M. et al. SOD1 mutations disrupt redox-sensitive Rac
accumulation in chronic traumatic encephalopathy. Acta Neuropathol. regulation of NADPH oxidase in a familial ALS model. J. Clin. Invest. 118,
Commun. 4, 112 (2016). 659–670 (2008).
54. Yoshiyama, Y. et al. Synapse loss and microglial activation precede tangles 85. Zhao, W. et al. Extracellular mutant SOD1 induces microglial-mediated
in a P301S tauopathy mouse model. Neuron 53, 337–351 (2007). motoneuron injury. Glia 58, 231–243 (2010).
55. Bhaskar, K. et al. Regulation of tau pathology by the microglial fractalkine 86. O’Rourke, J. G. et al. C9orf72 BAC transgenic mice display typical
receptor. Neuron 68, 19–31 (2010). pathologic features of ALS/FTD. Neuron 88, 892–901 (2015).
56. Nash, K. R. et al. Fractalkine overexpression suppresses tau pathology in a 87. O’Rourke, J. G. et al. C9orf72 is required for proper macrophage and
mouse model of tauopathy. Neurobiol. Aging 34, 1540–1548 (2013). microglial function in mice. Science 351, 1324–1329 (2016).
57. Bemiller, S. M. et al. TREM2 deficiency exacerbates tau pathology through 88. Spiller, K. J. et al. Microglia-mediated recovery from ALS-relevant motor
dysregulated kinase signaling in a mouse model of tauopathy. Mol. neuron degeneration in a mouse model of TDP-43 proteinopathy. Nat.
Neurodegener. 12, 74 (2017). Neurosci. 21, 329–340 (2018).
58. Deng, H., Wang, P. & Jankovic, J. The genetics of Parkinson disease. Ageing 89. Paolicelli, R. C. et al. TDP-43 depletion in microglia promotes amyloid
Res. Rev. 42, 72–85 (2018). clearance but also induces synapse loss. Neuron 95, 297–308.e6 (2017).
59. Dickson, D. W. Neuropathology of Parkinson disease. Parkinsonism Relat. 90. Iaccarino, L. et al. An in vivo 11C-(R)-PK11195 PET and in vitro pathology
Disord. 46(Suppl 1), S30–S33 (2018). study of microglia activation in Creutzfeldt-Jakob disease. Mol. Neurobiol.
60. McGeer, P. L., Itagaki, S., Boyes, B. E. & McGeer, E. G. Reactive microglia 55, 2856–2868 (2018).
are positive for HLA-DR in the substantia nigra of Parkinson’s and 91. Yamasaki, T., Suzuki, A., Hasebe, R. & Horiuchi, M. Flow cytometric
Alzheimer’s disease brains. Neurology 38, 1285–1291 (1988). detection of PrPSc in neurons and glial cells from prion-infected mouse
61. Gerhard, A. et al. In vivo imaging of microglial activation with [11C] brains. J. Virol. 92, e01457–17 (2017).
(R)-PK11195 PET in idiopathic Parkinson’s disease. Neurobiol. Dis. 21, 92. Falsig, J. et al. A versatile prion replication assay in organotypic brain slices.
404–412 (2006). Nat. Neurosci. 11, 109–117 (2008).
62. Halliday, G. M. & Stevens, C. H. Glia: initiators and progressors of 93. Sorce, S. et al. The role of the NADPH oxidase NOX2 in prion
pathology in Parkinson’s disease. Mov. Disord. 26, 6–17 (2011). pathogenesis. PLoS Pathog. 10, e1004531 (2014).
63. Croisier, E., Moran, L. B., Dexter, D. T., Pearce, R. K. & Graeber, M. B. 94. Aguzzi, A. & Zhu, C. Microglia in prion diseases. J. Clin. Invest. 127,
Microglial inflammation in the parkinsonian substantia nigra: relationship 3230–3239 (2017).
to alpha-synuclein deposition. J. Neuroinflammation 2, 14 (2005). 95. Hafner-Bratkovič, I., Benčina, M., Fitzgerald, K. A., Golenbock, D. & Jerala,
64. Su, X. et al. Synuclein activates microglia in a model of Parkinson’s disease. R. NLRP3 inflammasome activation in macrophage cell lines by prion
Neurobiol. Aging 29, 1690–1701 (2008). protein fibrils as the source of IL-1β​and neuronal toxicity. Cell. Mol. Life
65. Kim, C. et al. Neuron-released oligomeric α​-synuclein is an endogenous Sci. 69, 4215–4228 (2012).
agonist of TLR2 for paracrine activation of microglia. Nat. Commun. 4, 96. Xie, W. L. et al. Abnormal activation of microglia accompanied with
1562 (2013). disrupted CX3CR1/CX3CL1 pathway in the brains of the hamsters infected
66. Kuhlmann, T. et al. An updated histological classification system for with scrapie agent 263K. J. Mol. Neurosci. 51, 919–932 (2013).
multiple sclerosis lesions. Acta Neuropathol. 133, 13–24 (2017). 97. Hughes, M. M., Field, R. H., Perry, V. H., Murray, C. L. & Cunningham, C.
67. Yamasaki, R. et al. Differential roles of microglia and monocytes in the Microglia in the degenerating brain are capable of phagocytosis of beads
inflamed central nervous system. J. Exp. Med. 211, 1533–1549 (2014). and of apoptotic cells, but do not efficiently remove PrPSc, even upon LPS
68. Zrzavy, T. et al. Loss of ‘homeostatic’ microglia and patterns of their stimulation. Glia 58, 2017–2030 (2010).
activation in active multiple sclerosis. Brain 140, 1900–1913 (2017). 98. Sakai, K. et al. Absence of CD14 delays progression of prion
69. Ghosh, R. & Tabrizi, S. J. Huntington disease. Handb. Clin. Neurol. 147, diseases accompanied by increased microglial activation. J. Virol. 87,
255–278 (2018). 13433–13445 (2013).
70. Tai, Y. F. et al. Microglial activation in presymptomatic Huntington’s disease 99. Kouadir, M. et al. CD36 participates in PrP(106-126)-induced activation of
gene carriers. Brain 130, 1759–1766 (2007). microglia. PLoS One 7, e30756 (2012).
71. Pavese, N. et al. Microglial activation correlates with severity in Huntington 100. Simonian, N. A. & Coyle, J. T. Oxidative stress in neurodegenerative
disease: a clinical and PET study. Neurology 66, 1638–1643 (2006). diseases. Annu. Rev. Pharmacol. Toxicol. 36, 83–106 (1996).
72. Sapp, E. et al. Early and progressive accumulation of reactive microglia in the 101. Brown, G. C. & Vilalta, A. How microglia kill neurons. Brain Res.
Huntington disease brain. J. Neuropathol. Exp. Neurol. 60, 161–172 (2001). 1628(Pt B), 288–297 (2015).
73. Crotti, A. et al. Mutant huntingtin promotes autonomous microglia 102. Maezawa, I. & Jin, L. W. Rett syndrome microglia damage dendrites
activation via myeloid lineage-determining factors. Nat. Neurosci. 17, and synapses by the elevated release of glutamate. J. Neurosci. 30,
513–521 (2014). 5346–5356 (2010).
74. Lall, D. & Baloh, R. H. Microglia and C9orf72 in neuroinflammation and 103. Gan, L. et al. Identification of cathepsin B as a mediator of neuronal death
ALS and frontotemporal dementia. J. Clin. Invest. 127, 3250–3258 (2017). induced by Abeta-activated microglial cells using a functional genomics
75. Henkel, J. S. et al. Presence of dendritic cells, MCP-1, and activated approach. J. Biol. Chem. 279, 5565–5572 (2004).
microglia/macrophages in amyotrophic lateral sclerosis spinal cord tissue. 104. Leonardo, C. C., Hall, A. A., Collier, L. A., Gottschall, P. E. & Pennypacker,
Ann. Neurol. 55, 221–235 (2004). K. R. Inhibition of gelatinase activity reduces neural injury in an ex vivo
76. Turner, M. R. et al. Evidence of widespread cerebral microglial activation in model of hypoxia-ischemia. Neuroscience 160, 755–766 (2009).
amyotrophic lateral sclerosis: an [11C](R)-PK11195 positron emission 105. Hickman, S. E. & El Khoury, J. TREM2 and the neuroimmunology of
tomography study. Neurobiol. Dis. 15, 601–609 (2004). Alzheimer’s disease. Biochem. Pharmacol. 88, 495–498 (2014).

1368 Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience


FOCUS | Review
| Article
NATure NeurosCIenCe FOCUShttps://doi.org/10.1038/s41593-018-0242-x
Review Article
106. Yeh, F. L., Wang, Y., Tom, I., Gonzalez, L. C. & Sheng, M. TREM2 Binds to 132. Calvo, A. et al. Common polymorphisms of chemokine (C-X3-C motif)
apolipoproteins, including APOE and CLU/APOJ, and thereby facilitates receptor 1 gene modify amyotrophic lateral sclerosis outcome: a population-
uptake of amyloid-beta by microglia. Neuron 91, 328–340 (2016). based study. Muscle Nerve 57, 212–216 (2018).
107. Wang, Y. et al. TREM2 lipid sensing sustains the microglial response in an 133. Grizenkova, J., Akhtar, S., Brandner, S., Collinge, J. & Lloyd, S. E. Microglial
Alzheimer’s disease model. Cell 160, 1061–1071 (2015). Cx3cr1 knockout reduces prion disease incubation time in mice. BMC
108. Zhao, Y. et al. TREM2 is a receptor for β​-amyloid that mediates microglial Neurosci. 15, 44 (2014).
function. Neuron 97, 1023–1031.e7 (2018). 134. Striebel, J. F., Race, B., Carroll, J. A., Phillips, K. & Chesebro, B. Knockout
109. Suárez-Calvet, M. et al. Early changes in CSF sTREM2 in dominantly of fractalkine receptor Cx3cr1 does not alter disease or microglial activation
inherited Alzheimer’s disease occur after amyloid deposition and neuronal in prion-infected mice. J. Gen. Virol. 97, 1481–1487 (2016).
injury. Sci. Transl. Med. 8, 369ra178 (2016). 135. PrabhuDas, M. R. et al. A consensus definitive classification of scavenger
110. Takahashi, K., Rochford, C. D. & Neumann, H. Clearance of apoptotic receptors and their roles in health and disease. J. Immunol. 198,
neurons without inflammation by microglial triggering receptor expressed 3775–3789 (2017).
on myeloid cells-2. J. Exp. Med. 201, 647–657 (2005). 136. Cornejo, F. et al. Scavenger receptor-A deficiency impairs immune response
111. Wang, Y. et al. TREM2-mediated early microglial response limits diffusion of microglia and astrocytes potentiating Alzheimer’s disease
and toxicity of amyloid plaques. J. Exp. Med. 213, 667–675 (2016). pathophysiology. Brain Behav. Immun. 69, 336–350 (2018).
112. Ulland, T. K. et al. TREM2 maintains microglial metabolic fitness in 137. Šerý, O. et al. CD36 gene polymorphism is associated with Alzheimer’s
Alzheimer’s disease. Cell 170, 649–663.e13 (2017). disease. Biochimie 135, 46–53 (2017).
113. Condello, C., Yuan, P., Schain, A. & Grutzendler, J. Microglia constitute a 138. Wilkinson, K., Boyd, J. D., Glicksman, M., Moore, K. J. & El Khoury, J. A
barrier that prevents neurotoxic protofibrillar Aβ​42 hotspots around high content drug screen identifies ursolic acid as an inhibitor of amyloid
plaques. Nat. Commun. 6, 6176 (2015). beta protein interactions with its receptor CD36. J. Biol. Chem. 286,
114. Jay, T. R., von Saucken, V. E. & Landreth, G. E. TREM2 in 34914–34922 (2011).
neurodegenerative diseases. Mol. Neurodegener. 12, 56 (2017). 139. Origlia, N. et al. Microglial receptor for advanced glycation end product-
115. Song, W. M. et al. Humanized TREM2 mice reveal microglia-intrinsic dependent signal pathway drives beta-amyloid-induced synaptic depression
and -extrinsic effects of R47H polymorphism. J. Exp. Med. 215, and long-term depression impairment in entorhinal cortex. J. Neurosci. 30,
745–760 (2018). 11414–11425 (2010).
116. Paloneva, J. et al. DAP12/TREM2 deficiency results in impaired osteoclast 140. Vodopivec, I. et al. RAGE does not affect amyloid pathology in transgenic
differentiation and osteoporotic features. J. Exp. Med. 198, 669–675 (2003). ArcAbeta mice. Neurodegener. Dis. 6, 270–280 (2009).
117. Satoh, J. et al. Immunohistochemical characterization of microglia in 141. Chitramuthu, B. P., Bennett, H. P. J. & Bateman, A. Progranulin: a new
Nasu-Hakola disease brains. Neuropathology 31, 363–375 (2011). avenue towards the understanding and treatment of neurodegenerative
118. Suárez-Calvet, M. et al. sTREM2 cerebrospinal fluid levels are a potential disease. Brain 140, 3081–3104 (2017).
biomarker for microglia activity in early-stage Alzheimer’s disease 142. Gong, Y. et al. Microglial dysfunction as a key pathological change in
and associate with neuronal injury markers. EMBO Mol. Med. 8, adrenomyeloneuropathy. Ann. Neurol. 82, 813–827 (2017).
466–476 (2016). 143. Chen, Y. et al. Association of progranulin polymorphism rs5848 with
119. Lue, L. F. et al. TREM2 protein expression changes correlate with neurodegenerative diseases: a meta-analysis. J. Neurol. 262, 814–822 (2015).
Alzheimer’s disease neurodegenerative pathologies in post-mortem 144. Minami, S. S. et al. Progranulin protects against amyloid β​ deposition
temporal cortices. Brain Pathol. 25, 469–480 (2015). and toxicity in Alzheimer’s disease mouse models. Nat. Med. 20,
120. Lill, C. M. et al. The role of TREM2 R47H as a risk factor for Alzheimer’s 1157–1164 (2014).
disease, frontotemporal lobar degeneration, amyotrophic lateral sclerosis, 145. Van Kampen, J. M., Baranowski, D. & Kay, D. G. Progranulin gene delivery
and Parkinson’s disease. Alzheimers Dement. 11, 1407–1416 (2015). protects dopaminergic neurons in a mouse model of Parkinson’s disease.
121. Cady, J. et al. TREM2 variant p.R47H as a risk factor for sporadic PLoS One 9, e97032 (2014).
amyotrophic lateral sclerosis. JAMA Neurol. 71, 449–453 (2014). 146. Sleegers, K. et al. Progranulin genetic variability contributes to amyotrophic
122. Jiang, T. et al. Silencing of TREM2 exacerbates tau pathology, lateral sclerosis. Neurology 71, 253–259 (2008).
neurodegenerative changes, and spatial learning deficits in P301S tau 147. Erny, D. et al. Host microbiota constantly control maturation and function
transgenic mice. Neurobiol. Aging 36, 3176–3186 (2015). of microglia in the CNS. Nat. Neurosci. 18, 965–977 (2015).
123. Leyns, C. E. G. et al. TREM2 deficiency attenuates neuroinflammation 148. Vogt, N. M. et al. Gut microbiome alterations in Alzheimer’s disease.
and protects against neurodegeneration in a mouse model of tauopathy. Sci. Rep. 7, 13537 (2017).
Proc. Natl. Acad. Sci. USA 114, 11524–11529 (2017). 149. Minter, M. R. et al. Antibiotic-induced perturbations in gut microbial
124. Hickman, S. E. & El Khoury, J. Mechanisms of mononuclear phagocyte diversity influences neuro-inflammation and amyloidosis in a murine
recruitment in Alzheimer’s disease. CNS Neurol. Disord. Drug Targets 9, model of Alzheimer’s disease. Sci. Rep. 6, 30028 (2016).
168–173 (2010). 150. Sampson, T. R. et al. Gut microbiota regulate motor deficits and
125. Zujovic, V., Schussler, N., Jourdain, D., Duverger, D. & Taupin, V. In vivo neuroinflammation in a model of Parkinson’s disease. Cell 167,
neutralization of endogenous brain fractalkine increases hippocampal 1469–1480.e12 (2016).
TNFalpha and 8-isoprostane production induced by intracerebroventricular
injection of LPS. J. Neuroimmunol. 115, 135–143 (2001).
126. Liu, Z., Condello, C., Schain, A., Harb, R. & Grutzendler, J. CX3CR1 in
Acknowledgements
This work was supported by NIH grant RF1 AG051506 to J.E.K.
microglia regulates brain amyloid deposition through selective protofibrillar
amyloid-β​ phagocytosis. J. Neurosci. 30, 17091–17101 (2010).
127. Morganti, J. M. et al. The soluble isoform of CX3CL1 is necessary for Competing interests
neuroprotection in a mouse model of Parkinson’s disease. J. Neurosci. 32, The authors declare no competing interests.
14592–14601 (2012).
128. Cardona, A. E. et al. Control of microglial neurotoxicity by the fractalkine
receptor. Nat. Neurosci. 9, 917–924 (2006). Additional information
129. Thome, A. D., Standaert, D. G. & Harms, A. S. Fractalkine signaling Supplementary information is available for this paper at https://doi.org/10.1038/
regulates the inflammatory response in an α​-synuclein model of Parkinson s41593-018-0242-x.
disease. PLoS One 10, e0140566 (2015). Reprints and permissions information is available at www.nature.com/reprints.
130. Martin, I. et al. Ribosomal protein s15 phosphorylation mediates LRRK2
neurodegeneration in Parkinson’s disease. Cell 157, 472–485 (2014). Correspondence should be addressed to J.E.
131. Lopez-Lopez, A. et al. CX3CR1 is a modifying gene of survival and Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
progression in amyotrophic lateral sclerosis. PLoS One 9, e96528 (2014). published maps and institutional affiliations.

Nature Neuroscience | VOL 21 | OCTOBER 2018 | 1359–1369 | www.nature.com/natureneuroscience 1369

You might also like