You are on page 1of 14

ReVIeWS

I­mmune cell regulation of glia during


CNS injury and disease
Andrew D. Greenhalgh   1 ✉, Sam David2 and F. Chris  Bennett   3
Abstract | Glial cells are abundant in the CNS and are essential for brain development and
homeostasis. These cells also regulate tissue recovery after injury and their dysfunction is a
possible contributing factor to neurodegenerative and psychiatric disease. Recent evidence
suggests that microglia, which are also the brain’s major resident immune cells, provide
disease-modifying regulation of the other major glial populations, namely astrocytes and
oligodendrocytes. In addition, peripheral immune cells entering the CNS after injury and in
disease may directly affect microglial, astrocyte and oligodendrocyte function, suggesting an
integrated network of immune cell–glial cell communication.

Macrophages
Despite the neuron’s fundamental importance, it has and present an unrealized opportunity to create new
Phagocytic immune cells long been known that the vertebrate CNS cannot ade- therapeutics.
involved in homeostasis and quately function without the brain’s non-​neuronal cells, In this Review, we briefly describe the major glial sub-
inflammation. They may be which mainly consist of glia1. Indeed, more than half of types in the brain and review a growing body of research
tissue resident or recruited
the cells in the CNS are glia2. A traditional definition that demonstrates the direct engagement of the immune
from circulating monocyte
populations to sites of
suggests that glia are distinct from neurons in their lack system with glia. In addition to glial and immune cells,
inflammation. of axons, dendrites and synaptic junctions3; however, other non-​neuronal cells in the CNS include endothelial
many decades of research have refined this definition cells, smooth muscle cells and pericytes that are asso-
by uncovering the unique and highly specialized mor- ciated with and form the cerebrovasculature18, menin-
phology and functions of different types of glial cell. geal fibroblasts that can produce extracellular matrix
Furthermore, recent advances in technologies for cell after injury19, and progenitor cells expressing the pro-
isolation, RNA sequencing and high-​d imensional, teoglycan NG2 that can give rise to glia and neurons20.
single-​cell analyses have allowed researchers to probe The effect of immune cell-​derived inflammation on the
the details of the brain’s resident cell populations and cerebrovasculature as a whole has been well investigated
revealed their heterogeneity within and between dif- and reviewed elsewhere21,22, whereas little is known of
ferent CNS structures4–8. As a result, we now have an direct immune cell modulation of NG2-expressing cells.
enormous wealth of data on the type of cells present in Therefore, these non-​neuronal cells are not considered
the brain and what they look like. further herein.
The subtypes of mature CNS glia are astrocytes, oli-
1
INRA, Laboratoire Nutrition
godendrocytes and microglia. As the brain’s resident CNS glial cells
et Neurobiologie Intégrée, macrophages, microglia are classified as both glia and Astrocytes. Astrocytes are found throughout the CNS,
University of Bordeaux, immune cells, making the immune system an omni- with each cell containing numerous stellate processes
Bordeaux, France. present cellular feature of the brain. Beyond microglia, forming a well‐delineated bushy territory23. The complex
2
Centre for Research in recent work shows that the brain and its barriers host a fine morphology of these processes results in intimate
Neuroscience, The Research surprisingly diverse cohort of other immune cells that contacts with synapses, blood vessels and other glial
Institute of the McGill
University Health Center,
are more traditionally associated with the periphery9–12 cells. Indeed, astrocytes are required for the formation
Montreal, Quebec, Canada. (Fig. 1). During injury or disease, these resident immune and function of synapses24,25. They also support neurons
3
Department of Psychiatry, cells are joined by their infiltrating counterparts from through tight metabolic coupling26,27, modulate neuronal
University of Pennsylvania the circulation12,13. Brain immune cells, whether resident excitability and plasticity28, and directly affect local syn-
Perelman School of Medicine or actively recruited, can directly interact with neurons; aptic function through rapid neurotransmitter uptake29
and the Children’s Hospital we refer the reader to multiple excellent reviews of this and extracellular potassium buffering30,31. Astrocytes are
of Philadelphia,
Philadelphia, PA, USA.
topic14–17. Less widely appreciated, however, is the fact a critical component of the ‘neurovascular unit’, contain-
✉e-​mail: adgreenhalgh@ that microglia and infiltrating peripheral immune cells ing cerebral blood vessels, pericytes and neurons, which,
gmail.com can also modulate brain function via their interactions among other things, helps maintain cerebral blood flow
https://doi.org/10.1038/ with glia. These functional interactions appear to have under physiological and pathological conditions32. The
s41583-020-0263-9 great bearing on the pathogenesis of many diseases consequences of astrocyte function are seen in many

Nature Reviews | Neuroscience


Reviews

a
Mature CNS-resident glia

Astrocytes Microglia CNS-resident Monocytes Monocyte- Dendritic Neutrophils Lymphocytes Natural Innate
(parenchymal macrophages derived cells killer cells lymphoid cells
macrophages) (perivascular, macrophages B cells
subdural and
meningeal)

T cells Eosinophils Mast cells

Established during Reside in the borders of the


embryonic development and CNS and can be recruited from
form stable populations the circulation during injury
Oligodendrocytes
(derived from OPCs)
CNS-resident immune cells

Activated
microglial cell

Fig. 1 | cNs resident cell types. a | Astrocytes, oligodendrocytes and microglia represent the major glial populations
of the CNS. Microglia are also a stable population of CNS-​resident immune cells, as are the populations of perivascular,
subdural and meningeal macrophages that are established during embryonic development. Other immune cells reside
in the borders of the CNS and can also be recruited from the circulation during disease and injury. This allows the direct
signalling of monocytes, monocyte-​derived macrophages, dendritic cells, neutrophils and lymphocytes (T cells and B cells)
to CNS-​resident glia. Other immune cell populations found in the borders of the CNS also include natural killer cells, innate
lymphoid cells, eosinophils and mast cells12, but little is known of their modulation of glia. b | Schematic illustrating a possible
integrated network of glia–immune cell interactions: solid arrows represent the known immune cell to glia communication
pathways that are the subject of the present review , dashed arrows represent direct interaction pathways that are either
not covered in this review or are yet to be investigated. OPC, oligodendrocyte progenitor cell.

domains, from memory consolidation to the generation which maximizes axonal conduction velocity39. Myelin
of circadian rhythms33,34. Astrocytes also have the capac- sheaths are in fact large extensions of the oligodendro-
ity to propagate inflammation, which can contribute to cyte cell membrane and are highly complex structures40.
disease development and progression35. Following CNS Myelinating oligodendrocytes are derived from oligo-
injury and during CNS disease, astrocytes form func- dendrocyte progenitor cells (OPCs) during embryo-
tional barriers that restrict the entry of inflammatory genesis and in the early stages of postnatal life41. An
cells into CNS parenchyma36 and, although long thought undifferentiated pool of OPCs is maintained in the
to restrict neuronal regeneration37, may in some cases adult CNS42, which can give rise to myelinating oligo-
provide support to regenerating neurons36,38. Due to the dendrocytes throughout adulthood39. This is important
critical functions of astrocytes in health, disease and because the proliferation and differentiation of OPCs
injury, the influence of the immune system on astrocytes into myelinating oligodendrocytes increase after CNS
is of burgeoning interest. injury and disease and may be crucial for the recovery
phases of such pathologies39. In addition, recent evidence
Oligodendrocytes. Mature oligodendrocytes are found suggests that oligodendrocytes may also have immune-​
throughout the grey and white matter in the CNS, related functions43, modulate network activity through
where their processes wrap around neuronal axons direct connections with the neuronal soma44 and have
and are responsible for the production of myelin, important interactions with the vasculature45. Therefore,

www.nature.com/nrn
Reviews

immune cell regulation of oligodendrocytes is likely to CNS resident immune cells


have wide ranging consequences in health and disease. Microglia are the numerically dominant CNS resi-
dent immune cell type under steady-​state conditions12.
Microglia. Microglia are found in all regions of the However, single-​cell RNA sequencing, mass cytome-
CNS and can be broadly defined as both glia and try and fate-​mapping techniques have revealed other
immune cells; more specifically, they can be defined immune cell populations in the CNS, largely residing in
Genome-​wide association as process-​bearing, highly ramified myeloid cells and the interfaces between the brain and the periphery7,11,12,59
studies tissue-​resident macrophages46. They are the principle (Fig. 1). One such interface is the brain’s lymphatic system,
Studies that integrate whole immune cell of the brain and have, for many decades, which is quickly becoming a new frontier in neuro­
genomes across populations to
been known to be involved in the brain’s response to immunological research60. The cerebrospinal fluid (CSF)
identify genetic variants in
individuals and associate those injury47. Interest in these cells has grown in recent years also contains a variety of immune cells61,62 and several
variants with a trait. due to four distinct findings. First, it was shown that reports highlight distinct adaptive immune cell popu-
the cellular processes of microglia are highly dynamic lations in the meninges and choroid plexus62,63. In the
Lymphatic system in the healthy brain and respond rapidly to injury48,49. steady state, over 20% of the brain’s immune cell com-
A vascular network that drains
fluid from the extracellular
Second, genome-​wide association studies have implicated partment is comprised of cells other than microglia12.
space and traffics this and mutations in many genes expressed selectively or at Half of these cells are described as barrier-​associated
immune cells to lymphoid high levels in microglia as risk factors for neurodegen- macrophages, with distinct profiles from parenchy-
organs and the systemic erative disease50,51. Third, microglia have been shown to mal microglia11,59. The remaining brain immune cells
circulation.
perform critical roles during neuronal circuit develop- in the steady state consist of neutrophils, monocytes,
Adaptive immune cell ment, including a role in complement-​mediated synapse dendritic cells, innate lymphoid cells and lymphocyte sub-
A T cell or B cell that can elimination52,53. Finally, unlike other tissue-​resident macro­ sets, which differ in abundance and marker expression
acquire and maintain specific phages, microglia have been shown to be derived from to those circulating in the blood from which they are
knowledge of non-​ the embryonic yolk sac, suggesting early divergence thought to be derived12,59,64. The brain and its barriers
self-pathogens (or self-​antigens
during disease) to mount a
from the postnatal blood system during development are replete with innate and adaptive immune cells. How
specific immune response. and functional adaptation to the brain environment54,55. these cells interact with glia in health and disease is
Indeed, the introduction of a somatic mutation specif- discussed below.
Neutrophils ically in the erythro-​myeloid progenitor lineage from
Innate immune cells that are
which microglia are derived can drive late-​onset neuro­ Microglial regulation of other glia
often the first recruits to sites
of inflammation, where they degeneration in mice, suggesting an intrinsic role of these Microglial regulation of astrocytes. Interest in the inter-
release antimicrobial agents, tissue-​specific macrophages in adult disease56. action of glial cell subtypes is not new65. For example,
enzymes, nitrogen oxides and Although still poorly understood, it has become clear in vitro studies have long suggested that astrocytes are
other proteins. that microglia are more than quiescent sentinels that important for microglial homeostasis66,67. The tech-
Monocytes
stand ready to mount an immune response57. Instead, niques used to isolate and culture astrocytes and microg­
Cells that develop in the bone they are highly reactive to changes in the brain, actively lia are being continuously refined to ensure that the
marrow, are released into the producing signalling molecules to facilitate homeo­ cultured cells provide the best representation of the cells
circulation and can be stasis or to contribute to disease14–17. The study of ‘reac- in vivo; however, it should be noted that it is likely that
recruited during inflammation.
tive’ microglia has traditionally relied on the use of their the recreation of in vivo conditions in a dish is simply
Monocytes give rise to
monocyte-​derived cells such as morphological phenotype to indicate their distinct not possible68–74. For example, gene expression is rapidly
macrophages (but are often functional states, wherein the classification of ‘quiescent’ altered when microglia, astrocytes or oligodendrocytes
functionally distinct from versus ‘reactive’ is usually signified by a reduction in are placed in culture68–75. In addition, there are critical
tissue-​resident macrophage ramification57. However, microglial morphology is more species differences in disease-​related genes, such as
populations).
nuanced and highly dynamic and, although changes in APOE, and microglial immune responses can differ
Dendritic cells morphology are likely to reflect changes in the cell state, between mouse and human cells when in culture76.
Antigen-​presenting cells that it does not yet have robustly described molecular under- However, these caveats should not preclude the use of
exert immune surveillance for pinnings in vivo. New technologies that can capture fine these techniques and data; indeed, mixed culture, trans-​
exogenous and endogenous
morphology prior to the extraction of high-​dimensional well culture and conditioned medium culture experi-
antigens for the activation of
naive T cells giving rise to gene or protein signatures have the potential to create a ments have shed light on many mechanisms by which
various immunological lexicon that will allows us to better describe dynamic microglia modulate astrocyte function and we now have
responses. microglial states; however, with the exception of clear multiple methodologies for the further study of these
phagocytic activity, where phagosomes can be clearly immune cell–glial cell interactions (Box 1).
Innate lymphoid cells
Immune cells that belong to
visualized58, it is currently impossible to ascribe function The investigation of inflammatory processes in vitro,
the lymphoid lineage but do to morphology alone. Here, we define microglia ‘activa- with subsequent in vivo validation, has provided strong
not express antigen-​specific tion’ or ‘reactivity’ as any induced change in morphology or evidence for a direct microglial modulation of astro-
receptors and are therefore gene/protein expression from the homeostatic state, cytes. In vitro, combinations of the microglia-​derived
involved in innate immune
which will be further qualified by functional or molecular cytokines IL-1β, IL-6 and tumour necrosis factor (TNF)
regulation of homeostasis and
inflammation. information if described. drive the inhibition of astrocytic gap junctions, neces-
Microglia are both first responders and long-​lived sary for normal glial function77,78, and increase astrocytic
Lymphocyte resident immune cells of the CNS, containing hetero- glucose uptake while restricting its intercellular traffick-
A cell that is derived from a geneous and plastic populations, which places them in ing79. The activation of cultured astrocytes during man-
haematopoietic stem cell.
Lymphocytes ultimately form
a unique position to shape disease and injury outcomes. ganese neurotoxicity is contingent on nuclear factor-​κB
the cells of the adaptive Therefore, microglial regulation of other non-​neuronal (NF-​κ B)-dependent microglial release of TNF 80.
immune system. cell types is of great interest. Additionally, irradiation-​induced astrogliosis (defined as

Nature Reviews | Neuroscience


Reviews

the molecular, cellular and functional changes in astro- microglia release TNF, IL-1α and the complement com-
cytes that occur in response to an insult)81 only occurs ponent C1q, and microglia-​conditioned medium that
in vitro in the presence of microglia and requires their contains these factors induces the activation of cul-
release of IL-1β and TNF82,83. Finally, activated cultured tured astrocytes, characterized by the upregulation of

Box 1 | Methodological considerations for studying immune cell–glial cell interactions


choice of model
Glia–immune cell interactions involve cells of different lineages, across regionally complex CNs tissues in which the
immune cells are often invaders and not residents. these layered and dynamic processes are not readily modelled
in vitro224. in addition, microglia, astrocytes and oligodendrocytes undergo dramatic transcriptional changes during
cell culture, which may compromise their in vivo relevance68–75. Despite this, reduced in vitro models are necessary to
answer many experimental questions and can open important lines of investigation for further validation225. In vivo
experimentation may raise ethical issues but provides the complexity necessary to model human processes, though
consideration of species differences is paramount. Organoids and other ‘in vivo like’ models may provide a compromise
between these competing considerations, with the additional potential to use human cells220,226. recent advances show
that microglia, astrocytes and oligodendrocytes can be present in or engrafted to organoids221,227; however, these have
not yet been extensively used to study glia–immune cell interactions.
identifying and isolating cells
Glia subsets are readily distinguishable from one another; however, different types of immune cells express overlapping
sets of genetic or protein markers and exhibit within-​cell-type heterogeneity, making it challenging to identify intended
populations for study. a major advance was the identification of a set of microglia ‘signature genes’, allowing for the
development of ‘microglia-​specific’ genetic tools and reagents70,102,111,228,229. it must be noted, however, that microglia
signature gene expression is often unstable in disease states, presenting a problem for cell identification, although this may
be addressed by genetic fate labelling218. isolation of specific CNs glial and immune cell populations from the brain can also
dramatically affect gene expression in those cells216,230,231. therefore, it is critical to optimize cell isolation strategies and, in
any presentation of the work using these techniques, to describe the methods used in detail. In the analysis of isolated cells,
high-​dimensional, single-​cell approaches (such as single-​cell RNA sequencing, mass cytometry and spatial transcriptomics)
are now used to avoid prospective cell purification and therefore reduce the bias of targeting predefined lineages7,12,216.
in addition, new analysis methods can identify ligand–receptor interaction pairs within and between cell lineages,
thereby identifying cell–cell communication in single-​cell RNA sequencing datasets223,232. However, removing cells from
their environment risks preferentially selecting those that survive isolation procedures and therefore requires further
systematic study.
Targeting or modifying cells
Genetic targeting and modification (most commonly with the Cre–lox system) provides temporal control of gene
modification in specific cell types but is dependent on the existence of specific and efficient Cre lines102,233,234. the use of
viral tools (such as adeno-​associated viruses) as vectors to transfer genes to areas of interest can increase the feasibility
and provide anatomical specificity. Currently, adeno-​associated viruses have been developed to target astrocytes and
oligodendrocytes but it remains unfeasible to target microglia in vivo using this method235. an alternative method to target
cell–cell communication is through the depletion of one cell type. Genetic strategies (such as mice expressing diphtheria
toxin receptor (DTR) or herpes simplex virus thymidine kinase (HSVTK) in target cells) or pharmacologic strategies (such as
inhibition of CSF1R) can be used to remove microglia from the CNS to varying degrees236–240. recently, deletion of the
fms-​intronic regulatory element (FIRE) of the mammalian Csf1r locus was shown to result in a complete lack of microglia
with no overt developmental abnormalities241, overcoming an issue observed in Csf1r−/− mice74. similarly, immunodeficient
mice, such as severe combined immunodeficient (SCID) mice or RAG1-deficient mice242, or more specific peripheral
immune cell depletion, such as using Ly6G-​targeting antibodies to deplete neutrophils in mice243, offer a practical but
non-​specific method to study the impacts of immune function on glia.

Problem Approach Features


Choice of model In vitro models Reductionist, tractable, variable validity
In vivo models Complex, challenging implementation, broader validity
‘In vivo like’ models (such as Technically challenging, improved validity
organoids or 3D cultures)
Identifying and Cell type-​specific markers Increasingly specific, requires valid detection tools (such as
isolating cells antibodies or transgenic Cre lines), often unstable in disease
Cell isolation Dramatically affects cell phenotypes, requires cell type-specific
optimization and customization to assay and cell type
Single-​cell approaches Reduces cell type marker bias, costly , technically challenging
Targeting/ Genetic targeting Allows temporally and cell type-​specific manipulations, limited
modifying cells by line availability
Viral recombination High specificity possible, immune stimulatory , cannot
manipulate microglia
Cell ablation Widely used for immune populations, practical, lack of
molecular control, off-​target/non-​specific effects

www.nature.com/nrn
Reviews

a cassette of reactivity genes, loss of normal astrocyte In addition, the effects of the depletion of microglia
functions (such as promotion of synapse formation) and after development suggest that microglia also maintain
gain of neurotoxic functions84. To summarize, in vitro oligod­endrocyte progenitor numbers in the healthy adult
studies show that communication between microglia mouse brain93, though this may be dependent on the
and astrocytes occurs largely through inflammatory microglial depletion strategy96. However, the influence
mediators, which is not surprising when the immune of microglia on oligodendrocytes is not always trophic:
functions of microglia are considered. the production of TNF, nitric oxide and complement
The effects of microglia-​driven inflammation may after microglial activation can facilitate oligodendro-
be detrimental and/or beneficial for the nervous system cyte cell death and phagocytosis by microglia in rat cells
depending on timing and context85–87, and microglia-​ in vitro97,98. In human cell cultures, activated microglia
driven astrocyte activation is not necessarily harmful. produce factors that have direct cytotoxic effects on neu-
For example, it has recently been shown that astro- ral progenitors, which results in a decreased number of
cytes provide a supportive substrate and are necessary oligodendrocyte lineage cells either directly or indirectly
for recovery after spinal cord injury38. Recent in vivo via astrocyte signalling to their progenitors99.
studies have therefore built upon in vitro data in mul- The studies described above suggest that microglia
tiple disease settings. To further assess the potentially support the normal development of oligodendrocytes
detrimental role of astrocytes activated by microglial but can be toxic in inflammatory environments. Due
TNF, IL-1α and C1q, astrocyte reactivity was prevented to the plasticity of microglia, studies have also assessed
through the generation of transgenic mice lacking all their trophic potential and ability to modulate oligoden-
three cytokines, inhibiting neuronal cell death follow- drocyte function and improve remyelination in diseases
ing optic nerve crush84. Neurotoxic astrocytes with a such as multiple sclerosis100. For example, conditioned
similar phenotype to those induced by microglia in medium from mouse microglia treated with IL-13 or
culture and in animal models are also observed in the IL-10 enhances oligodendrocyte survival and differen-
post-​mortem brain tissue of individuals with neuro- tiation in vitro101. The targeted depletion of microglia
degenerative diseases, including multiple sclerosis and expressing CD206 (a receptor that is upregulated in
Alzheimer disease84. Glucagon-​like peptide 1 receptor microglia after IL-13 or IL-10 treatment) in an in vivo
(GLP1R) agonist treatment has been shown to reduce mouse model of demyelination decreases oligoden-
pathology in two mouse models of Parkinson disease drocyte differentiation and delays remyelination101. In
by decreasing the microglia-​mediated activation of EAE, conditional depletion of TGFβ-​activated kinase 1
astrocytes to a neurotoxic activation state88. On the (TAK1) in microglia suppresses disease and reduces CNS
other hand, in a rodent traumatic brain injury (TBI) inflammation and axonal and myelin damage through
model, inhibition of microglial activity or microglial the cell-​autonomous inhibition of the NF-​κ B, Jun
ablation significantly disrupts the formation of the N-​terminal kinase (JNK) and extracellular-​signal-
astrocytic scar and increases brain injury, suggesting regulated kinase 1 (ERK1)/ERK2 pathways102. However,
that there is a neuroprotective microglia–astrocyte it is not known whether these beneficial effects are due to
interaction after TBI89. In experimental autoimmune the loss of direct signalling from microglia to oligoden-
encephalomyelitis (EAE) in mice, microglia-​derived drocytes. Oligodendrocyte differentiation and survival
transforming growth factor-​α (TGFα) reduces disease also requires signalling from the extracellular matrix103.
severity by signalling directly to astrocytes, whereas The adhesion G protein-​coupled receptor ADGRG1
microglia-​derived vascular endothelial growth factor-​β (also known as GPR56) is an evolutionarily con-
(VEGFβ) increases disease burden 90. Interestingly, served regulator of oligodendrocyte development104,105.
the microglial control of astrocyte function in EAE Transglutaminase 2 is produced by microglia signals
is, in turn, directly regulated by microbially derived to ADGRG1 on OPCs in the presence of the extra-
tryptophan metabolite signalling through the aryl cellular matrix and promotes OPC proliferation and
hydrocarbon receptor (AHR), suggesting that gut remyelination in animal models106.
microbiota composition may modulate both the The complex influence of microglia on oligodendro-
microglial response and disease severity90. The valida- cytes may also affect cognitive function. Humans that
tion of microglial control of astrocytes in various CNS receive chemotherapy have cognitive problems and per-
pathologies demonstrates the importance of microglial– sistent depletion of oligodendrocyte lineage cells107. In a
astrocyte relays (Fig. 2). However, these studies also mouse model, methotrexate chemotherapy was shown
highlight the complexity and context-​d ependent to induce persistent microglial activation, driving neuro-
nature of such cell-​to-cell interactions. logical and cognitive dysfunction, depletion and dysreg-
ulation of OPCs, deficits in myelination, and astrocyte
Microglial regulation of oligodendrocytes. Oligo­ activation107. Therefore, chemotherapy has been pro-
dendrocytes are essential for proper neuronal function posed to induce a ‘tri-​glial’ dysregulation, with microglia
and depend on signals derived from astrocytes and as the key detrimental modulator of both astrocytes and
microglia for their differentiation and survival91,92. In vivo oligodendrocytes.
studies have shown that, during brain develop­ment In summary, combined evidence from in vitro and
in mice, microglia support oligodendrocyte matura­ in vivo experiments reveals that microglia directly
tion and function93,94. Indeed, selective depletion of a and indirectly modulate oligodendrocyte survival and
developmentally regulated CD11c-​expressing microg­ function in development and can be either beneficial
lial subset leads to impaired primary myelination95. or detrimental in different pathological contexts (Fig. 2).

Nature Reviews | Neuroscience


Reviews

Steady-state
microglial cell
TNF, IL-1α and C1q
Activated (neurodegenerative
microglial cell disease/acute injury),
VEGFβ (EAE) or
Oligodendrocyte IL-1β, IL-6 and iNOS Astrocyte
(in vitro)
??

Microglial context Microglial context


• Methotrexate chemotherapy • Neurodegenerative disease
• High levels of iron • Acute injury
• Inflammation • EAE
Detrimental effect • Developmental loss of CD11c • Methotrexate Detrimental effects
on oligodendrocytes expression in microglia chemotherapy on astrocytes and
and outcome, and outcome
reduced myelination
IL-13, IL-10 and
transglutaminase 2 TGFα
de(re)myelination (EAE)

Microglial context
• TBI
Beneficial effects • EAE Beneficial effects
on oligodendrocytes • Demyelination or on astrocytes and
and outcome remyelination outcome

Fig. 2 | context-​dependent signalling of microglia to astrocytes and oligodendrocytes. Microglia respond to a variety
of injuries and CNS pathologies through the activation of different signalling pathways that allow them to communicate
to oligodendrocytes and astrocytes in different ways, depending on the context. Both beneficial and detrimental
microglial effects on oligodendrocytes and astrocytes in disease and injury have been described84,88,90,101,106,107. In the
schematic, three different contexts in which microglia may be activated and the downstream effects of their activation
on astrocytes and oligodendrocytes are illustrated. Microglial activation is indicated by the solid arrows and microglial
signalling to other glial cells is indicated by the dashed arrows. Microglial activation during chemotherapy , when iron
levels are high, during inflammation or as a result of a developmental loss of CD11c expression results in detrimental
effects on oligodendrocytes and reduced myelination, although the soluble mediators are unknown. Microglial activation
in traumatic brain injury (TBI), experimental autoimmune encephalomyelitis (EAE) or as a result of demyelination or
remyelination can have a beneficial effect on oligodendrocytes through IL-10, IL-13 or transglutaminase 2 release
and on astrocytes via transforming growth factor-​α (TGFα) signalling. In neurodegenerative disease, acute injury , EAE and
chemotherapy , microglia can drive astrocytes to a detrimental disease-​modifying activation state through a variety of
inflammatory mediators. These effects highlight the highly context-​dependent nature of microglia–oligodendrocyte and
microglia–astrocyte interactions. C1q, complement component C1q; iNOS, inducible nitric oxide synthase; TNF, tumour
necrosis factor ; VEGFβ, vascular endothelial growth factor-​β.

Peripheral immune regulation of glia transcriptional profiling techniques has revealed unique
Peripheral immune cell entry to the CNS. At sites of features of each cell type that have helped address this
CNS injury and disease, the recruitment of peripheral question8,70,110,111.
immune cells can also modulate tissue-​resident glia Many ‘microglia-​specific’ proteins (those that are
(Fig. 3). In healthy conditions, immune cell entry into highly enriched in microglia in comparison with mono-
the parenchyma is restricted by multiple specialized cytes and monocyte-​derived macrophages) have now
cellular structures such as the blood–brain barrier and been described in mouse and humans, and antibodies
the blood–CSF barrier32,108. During perturbations to the and genetic tools to investigate these cells are avail­
CNS, immune cells, such as neutrophils, monocyte-​ able13. Despite this, work is ongoing to determine the
derived macrophages and lymphocytes, can traffic stability of such markers in disease and injury contexts,
through intact or compromised barriers and enter as it has been shown that some highly specific micro-
the brain108. Distinguishing these peripheral immune glial markers, such as TMEM119 and P2RY12, are
cell populations from most glia is relatively straight- susceptible to downregulation during inflammation
forward due to their unique morphology and specific and disease111,113,114. Indeed, the concept of a ‘homeo-
lineage markers109. However, it has historically been static’ signature of microglia was coined to describe
difficult to distinguish resident microglia from infiltrat- how microglia-​defining genes are downregulated in
ing monocyte-​derived macrophages as they are both disease contexts115. Adding to the complexity of distin-
phagocytic myeloid cells that share many common fea- guishing microglia from other macrophages, several
tures8,70,110,111. Nevertheless, to evaluate the communica- groups have shown that blood monocytes can begin to
tion of infiltrating immune cells to resident glia, the two express many microglia signature genes after engraft-
cell types must be distinguishable. Fortunately, microglia ing in the brain, including TMEM119 (refs73,74,111,116).
and infiltrating monocyte-​derived macrophages are now These data show that some portion of microglial iden-
known to have different ontogeny112 and the advent of tity reflects the exposure to brain parenchymal signals

www.nature.com/nrn
Reviews

a b c d e f

Microglia-mediated injury response Acute


injury

Hours Days Weeks


Time after acute injury

a Neutrophil T cell b c
Parenchyma CNS barriers Blood vessel

Monocyte

CNS-resident
macrophage

Homeostatic Activated
microglial cell microglial cell

Homeostasis Acute injury triggers resident Neutrophil entry and


immune cell activation phagocytosis by microglia

d Monocyte-derived e f
macrophage

??

Monocyte-derived macrophage Lymphocyte entry and Treg cell Resolution of microglial activation
entry and signalling to microglia signalling to microglia

Fig. 3 | Peripheral immune cell infiltration and their dynamic interaction with microglia after cNs injury. After
various types of acute CNS injury , there is a stereotyped infiltration of immune cells over time. Current evidence predicts
that a direct interaction between peripheral immune cells infiltrating the CNS and microglia can reduce microglial
activation after injury71,160,198,199. This hypothesis is illustrated in the graph at the top of the figure. The y-​axis represents
the level of microglial response after acute injury , that is, the magnitude of each cell’s response and/or the number of
cells involved. The x-​axis represents time after the injury. a | Immune cell populations of the CNS in their homeostatic
state prior to injury. b | Microglia respond (are activated), within minutes, to acute injury and initiate a cascade of
inflammatory events that further activate other microglial cells and are involved in the recruitment of immune cells to
the injury site148. Little is known of specific responses of other resident immune cells in injury and disease. c | Neutrophils
are the first blood-​borne immune cells to respond to acute tissue damage, entering the CNS within hours of injury87,134,149.
Phagocytosis of neutrophils by microglia partly resolves the microglial inflammatory phenotype161. d | The first
monocyte-​derived macrophages arrive in the parenchyma after 2–3 days87,134,149 and can also suppress the activation
of microglia after acute injury71. e | Lymphocytes are recruited over a period of days to weeks after the original injury176–178.
Regulatory T (Treg) cell signalling to microglia suppresses proinflammatory cytokine and chemokine production and
further reduces microglial activation179–185. f | Weeks after the initial injury , infiltrating immune cells can still be found
at the injury site in reduced numbers, though it is unknown whether they continue to signal to microglia and affect
recovery in the long-​term.

Nature Reviews | Neuroscience


Reviews

in the cellular environment rather than cell origin. disease onset and progression114,143–147 and the direct
Nevertheless, blood-​derived macrophages engrafted in signalling of infiltrating macrophages to microglia has
the brain remain morphologically and transcriptom- not been studied.
ically different from microglia even after long-​term In the context of injury, microglial cells are activated
residence, suggesting functional non-​e quivalence. prior to the arrival of monocyte-​derived macrophages
The impact of developmental origin on brain macro­ and may actively participate in their recruitment148.
phage function thus remains unresolved and a topic of In acute injuries associated with frank tissue dam-
ongoing study. age, the first monocyte-​derived macrophages arrive
in the parenchyma after 2–3 days87,134,149. In models of
Monocyte-​derived macrophage regulation of CNS glia. SCI, EAE, TBI and stroke, the arrival of monocyte-​
Relatively few studies have addressed how monocyte-​ derived macrophages coincides with the downregul­
derived macrophages modulate resident microglia, ation of microglial inflammation and phagocytic
despite evidence of their infiltration into the CNS in functions71,140,141,150,151. In a bilaminar culture system
injury and disease. Early after injury and before the entry designed to assess direct macrophage–microglial
of peripheral macrophages, microglia have been shown interactions, it was shown that macrophages directly
to limit the initial damage caused by a variety of insults, suppress microglial phagocytosis 71. Interestingly,
including laser lesion in acute brain slices117, ischae- microglia were found to have the reciprocal effect on
mic stroke in animal models118, hippocampal lesions119 macrophages, acting to enhance macrophages’ phago-
and spinal cord trauma120. However, it also known that cytic activity71. Mathematical modelling of cytokine
microglia can, after injury, be driven to an inflamma- signalling indicates that the inflammatory cytokines
tory state that has detrimental effects on neuronal sur- IL-1β, TNF, IL-6 and IL-10 are key nodes in inflamma-
vival121,122 and their chronic activation is detrimental tory networks85,152 and it has been shown that macro­
to the brain123. Monocyte-​derived macrophage arrival phages suppress these nodes in both adult mouse and
into the CNS may have a role in modulating these microg­ human microglia in culture71. Macrophage-​derived
lial functions and thus affect outcome in CNS injury prostaglandin E2 mediates suppression of microglial
and disease. phagocytosis in the bilaminar culture system and can
During viral or bacterial CNS infections, inflam- release trophic factors to reduce microglia-​mediated
matory monocyte-​derived macrophages infiltrate the injury of cultured neurons153. In vivo, preventing infil-
brain124. It has been shown that infected peripherally tration of macrophages to the CNS increases microg­
derived macrophages can prime naive microglial cells lial activation after trauma in the brain and spinal
(that is, they can cause them to adopt a state such that cord71,154,155. Therefore, after acute injury, macrophages
subsequent stimuli produce an exaggerated response125), entering the CNS may provide a regulatory mechanism
causing microglia to express a range of inflamma- that controls acute and long-​term microglia-​mediated
tory mediators in vitro126–128. Similarly, peripherally inflammation.
derived macrophages that are directly stimulated with Fewer studies have investigated monocyte-​derived
interferon-​γ (IFNγ), a factor critical for innate and macrophage communication to astrocytes; how-
adaptive immunity against viral infection, produce ever, evidence suggests that soluble factors released
mediators that induce TNF and inducible nitric oxide by inflammatory macrophages induce a reactive gene
synthase expression in cultured microglia129. Such stud- expression pattern in these cells in vitro156. In culture,
ies show that activated peripheral macrophages can alter human monocytes potentiate the production and release
microglia function and provide a proof-​of-concept that of CCL2 from astrocytes, potentially providing a feed-
myeloid cells from the periphery may modulate myeloid forward mechanism to increase the further recruitment
cells of the CNS. of monocytes157. Indeed, human monocytes infected
There is a long history of investigation as to whether with Mycobacterium tuberculosis in culture can drive
monocyte-​derived macrophages are detrimental or ben- astrocytes to produce the matrix-​degrading enzyme
eficial to recovery after acute injury87,130,131. Studies in spi- MMP9, which is also localized to astrocytes in post
nal cord injury (SCI), TBI and stroke have shown both mortem tissue sections from patients with CNS tuber-
detrimental and beneficial effects of these cells132–134. The culosis158. In a brain stab wound injury model, reducing
route of entry (either through a compromised blood– monocyte-​derived macrophage invasion results in a
brain barrier or the blood–CSF barrier) has been pro- marked increase in astrocyte proliferation but reduced
posed to confer beneficial or detrimental properties scar formation due to less extracellular matrix deposi-
on infiltrating immune cells in models of stroke, SCI tion155. This results in a smaller lesion site and increased
and amyotrophic lateral sclerosis (ALS)135–139. In acute neuronal preservation155. These results suggest that, in
injury, recruited monocyte-​d erived macrophages both sterile and infectious CNS injury, monocytes drive
quickly become the numerically dominant myeloid cell astrocytes to detrimental phenotypes.
in lesions and their boundary zones; however, unlike Interestingly, these early studies suggest that infiltrat-
microglia, they are not thought to be long-​lived in the ing macrophages may have direct beneficial actions on
CNS140–142. In neurodegenerative disease, the slow pro- microglia but modulate astrocytes in a detrimental way.
gression of pathology makes the infiltration dynamics More studies are required to clearly define the macro­
of immune cells more difficult to quantify. In Alzheimer phage-microglia–astrocyte interaction and whether
disease, numerous studies report conflicting results oligo­dendrocytes play a defined role in such immune–glia
regarding the contribution of recruited monocytes to network signalling.

www.nature.com/nrn
Reviews

Neutrophil regulation of CNS glia. Neutrophils are signal to resident microglia during injury and disease to
the first blood-​b orne immune cells to respond to alter pathophysiology.
acute tissue damage, entering the CNS within hours In CNS disease, lymphocytes are best known for their
of injury87,134,149,159 (Fig. 3), and are also reported to be role in multiple sclerosis pathology. Multiple sclerosis is
present in the CNS in neurodegenerative conditions160. a chronic, demyelinating, inflammatory disease, a hall-
The clearance of neutrophils from the CNS by macro­ mark of which is infiltration of immune cells into the
phages is one of the key events in the resolution of CNS173. Multiple sclerosis stands alone as a major CNS
inflammation161,162. Neutrophil clearance is mediated disease for which multiple effective disease-​modifying
by the production of specialized pro-​resolution medi- drugs are available174. Several of these target lymphocyte
ators from macrophages that actively resolve inflam- function and/or entry to the CNS174,175. The investigation
mation. These bioactive lipids, derived from omega-3 of direct lymphocyte action on glia is still in its early
polyunsaturated fatty acids (including maresins, neuro- stages in the context of multiple sclerosis and lympho-
protectins and resolvins), promote recovery after CNS cyte function in the disease is the subject of several excel-
injury163,164. Intravital microscopy reveals that microglia lent reviews to which we refer the reader173–175. Due to
react to passing neutrophils, interacting with them and space limitations, we focus herein on lymphocyte–glia
engulfing them165 through mechanisms likely mediated interactions on acute injury and other diseases.
by specialized pro-​resolution mediators164. It has long Following acute injuries such as stroke or CNS
been established that the phagocytosis of neutrophils by trauma, lymphocytes are recruited after neutrophils and
macrophages, including microglia, results in a pheno- monocyte-​derived macrophages over a period that starts
typic and functional change of the phagocytic cells and days after the original injury and lasts for several further
is therefore a major form of microglia regulation122,166. In weeks176–178. Multiple lines of evidence from studies in
triple-​transgenic Alzheimer disease model mice, which animal models demonstrate that lymphocytes may play
develop age-​related progressive neuropathology, neutro­ a role in injury progression176,179. Furthermore, although
phil depletion results in less microglial activation160; the numbers of T cells and B cells in human post mortem
however, to our knowledge, evidence for direct signal- lesions are low180,181, it has been shown that lymphocytes
ling from neutrophils to microglia (rather than an effect accumulate in the meninges and that T cell numbers
mediated via phagocytosis) is lacking. increase in injured human CNS tissue over time130,180,182.
With regard to the communication of neutrophils In the chronic setting of neurodegenerative diseases,
to astrocytes, data from zebrafish indicate that, in the lymphocytes (mainly T cells) are present in post mortem
context of tumour progression, impaired neutrophil tissue taken from individuals with Alzheimer disease,
function reduces the proliferation of astrocytes and ALS and Parkinson disease, and have also been found
limits the progression of tumorigenesis167. However, in tissue derived from corresponding mouse models of
in human cell culture experiments, components of the these diseases183–190. Perhaps the most intensive interest
complement system released by neutrophils are reported in how lymphocytes affect CNS-​resident cells, at least in
to promote astroglial differentiation from neural stem commercial drug discovery, lies in the search for immuno­
cells and to promote migration168. Despite the limited therapies for neurodegenerative diseases, particularly
investigation into direct neutrophil signalling to resident in Alzheimer disease191. These therapies include active
glial cells, it is interesting to note that neutrophils pro- vaccines that aim to stimulate lymphocytes to produce
duce an array of cytokines and proteolytic enzymes that antibodies against amyloid-​β (one of the key proteins
can affect microglia, astrocytes and oligodendrocytes169. implicated in Alzheimer disease pathogenesis) or passive
Therefore, signalling between neutrophils and glial cells immunization through the administration of exogenous
can be expected to occur. The role of such signalling in antibodies. In both cases, the ultimate aim is to stimulate
disease and injury is an important area for future study. the clearance of amyloid-​β plaques through microglial
phagocytosis192. There is much hope surrounding these
Lymphocyte communication to CNS glia. Lymphocytes therapies but there have also been many setbacks, as
in the healthy brain are relatively scarce but their capac- reviewed elsewhere191. Here, we focus on evidence for
ity to secrete modulatory factors puts them in a good direct signalling of lymphocytes to resident CNS cells.
position to influence the resident cells of the CNS. Although critical players in the peripheral immune
Direct signalling between T cells and CNS cells has response, the role of B cells in CNS injury and neuro­
been shown in animal models to have a role in phys- degenerative disease is only beginning to be understood.
iological processes such as social behaviour15,170 and Mice treated with IL-10-producing B cells have reduced
learning, and it has been suggested that meningeal numbers of activated microglia after stroke, though it is
T cells increase astrocytic brain-​derived neurotrophic unclear whether B cell-​derived IL-10 acts directly on the
factor (BDNF) expression through the release of IL-4 microglia193. In models of traumatic SCI and stroke, B cells
(ref.171). It is therefore reasonable to assume that lympho- synthesize autoantibodies194,195, which are deposited
cytes recruited to the brain as part of an inflammatory together with C1q and activate cells bearing Fc recep-
response may have equally measurable effects on CNS tors, including microglia194. Removal of B cells in both
glia. A large body of literature exists describing IFNγ contexts improves recovery, though the direct detrimen-
and IL-4, which are known T cell derived cytokines, as tal effect of B cells remains to be determined. Indeed,
factors that respectively drive proinflammatory and modulation of B cell-​dependent pathological mecha-
anti-​inflammatory states in microglia in culture172. As nisms may be more important in the periphery than in
a result, there is mounting evidence that T cells directly the CNS as infections such as pneumonia are a major

Nature Reviews | Neuroscience


Reviews

complication of stroke and SCI and cause increased In support of such a beneficial T cell response, the expan­
morbidity and mortality196. sion of endogenous Treg cells in a mouse model of ALS
Many studies show that specific T cell subsets are significantly prolonged survival time and was associ-
either beneficial, such as regulatory T (Treg) cells, or det- ated with a marked suppression of glial cell immuno-
rimental, such as γδ T cells, after acute injury in animal reactivity209. In vitro evidence suggests that the cytokine
models62,197. Fewer studies have investigated the direct IL-33 may drive T cells to reduce astrocyte activation
signalling of T cells to resident glia. In in vivo stroke in ALS210. However, it is also known that oligodendro-
models, IL-10-producing Treg cells reduce the activa- cytes can release IL-33 in the injured spinal cord, which
tion of resident and invading immune cells, including can directly affect astrocytes and microglia211; therefore,
TNF-​producing microglia198,199. Interestingly, in cul- the actions of IL-33 on resident glia may also be derived
tures of cells isolated from the brains of mice undergo- from sources other than T cells.
ing beneficial immunosuppressive treatments, Treg cells In contrast to the evidence of beneficial actions in
directly suppress microglial proinflammatory cytokine models of ALS, in Parkinson disease models, it has
and chemokine production200,201. In animal models, been shown that T cells must be present for activation
Treg cells continue to accumulate in the brain 2 weeks of microglia and the resulting dopaminergic neuro­
after injury and appear to be important in long-​term degeneration after viral overexpression of α-​synuclein212.
neurological recovery as a result of their signalling to In vitro, T cells significantly reduced α-​synuclein phago-
astrocytes 182: they suppress neurotoxic astrogliosis cytosis by microglia213 and the absence of mature CD4-
by producing amphiregulin, a low-​affinity epidermal expressing T cells in vivo reduced microglial activation
growth factor receptor (EGFR) ligand182. In studies with in a MPTP model of the disease186.
less specific targeting of immune cells before SCI, such Lastly, in a model of chronic stress, mitochondrial fis-
as the complete removal of T cell and B cell subsets (in sion in peripheral CD4-expressing T cells drives anxiety-​
Rag2–/– mice), recovery from injury is improved and like behaviour through the production of the purine
there is reduced microglia/macrophage activation but xanthine214. Using almost all of the in vivo techniques
increased astrocyte numbers202, suggesting differential described in Box 1, it was shown that oligodendrocytes
effects of lymphocytes on the two glia subtypes. in the left amygdala were modulated by CD4-expressing
In human Alzheimer disease post mortem tissue, T cell-​derived xanthine through the adenosine A1 recep-
T cells are observed in close proximity to microglia, tor, and that this pathway is responsible for the anxiety-​
suggesting that direct signalling is possible183. However, like behaviour after chronic stress214. This highlights
it is still unclear whether the endogenous lymphocyte the power of the peripheral lymphocyte response to
response in Alzheimer disease is beneficial. When modulate behaviour through glia.
transgenic APP/PS1 mice (which accumulate amyloid-​ In summary, there is now clear evidence that lympho­
β deposits at a young age) are crossed with Rag2–/– mice, cytes can directly modulate resident glia and affect the
brain amyloid-​β pathology is decreased in association outcome of disease states in various animal models
with enhanced microglial activation and increased (Fig. 3). However, the nature of these interactions var-
phagocytosis of amyloid-​β peptide aggregates 203. ies not only between different disease models but also
Amyloid-​β-specific type 1 T helper CD4-expressing within models of the same disease.
cells, adoptively transferred to APP/PS1 mice, release
IFNγ, increase microglial activation and impair cogni- Future perspectives
tive function204. By contrast, immunodeficient Rag2–/– An abundance of in vitro and in vivo data now demon-
mice crossed with 5xfAD mice (which express human strates the powerful effects of immune signals on other
transgenes with a total of five Alzheimer disease-​linked non-​n euronal CNS cells throughout the lifespan.
mutations) exhibit a greater than twofold increase in Targeting these interactions therefore holds enormous
amyloid-​β pathology and increased microglia-​mediated promise for the treatment of neurological diseases of
neuroinflammation with reduced phagocytic capacity205. development, injury and ageing. Indeed, the preclinical
This suggests that the adaptive immune cell populations efficacy of this approach has already been demonstrated.
restrain Alzheimer disease pathology through microg­ For example, altering brain macrophage polarization fol-
lial actions205. These studies highlight important roles lowing demyelination appears to improve remyelination
for lymphocyte regulation of microglia in different dis- by supporting oligodendrocyte maturation, while microg­
ease models, yet it is clear that the various in vivo animal lial depletion resolves chemotherapy-​induced cognitive
models differ greatly in their recapitulation of the disease. decline101,107. However, immune signals appear to have
ALS is characterized by selective degeneration of highly context-​specific effects and simultaneous direct
motor neurons and, in some cases, is linked to mutations and indirect contributions from both peripheral and
in the superoxide dismutase (SOD1) gene206. In humans resident immune cells can modulate resident glia. To
and mutant Sod1-transgenic mice, loss of motor neurons add to this complexity, it is now apparent that immune
is accompanied by robust microglia and astrocyte acti- cells in brain barriers are bona fide resident populations,
vation207. Mutant Sod1 mice bred onto a T cell receptor distinct from parenchymal microglia and their counter-
β-​chain-deficient background, ablating αβ T cells, show parts in the periphery7. It is largely unknown how the
accelerated disease progression and reduced microglial heterogeneity of these cells contributes to CNS disease
reactivity, indicating that these T cells play an endo­g­ or injury, but their presence must be considered as a fac-
enous neuroprotective role in ALS by modulating a tor when developing therapies targeting immune–glia
beneficial inflammatory response to neuronal injury208. interactions.

www.nature.com/nrn
Reviews

This leads to a broad conceptual question — given the astrocytes restrain CNS inflammation through microg­
extensive data showing functionally relevant communi- lia219. Here, whether one set of responses precedes the
cation between glia and nearly all classes of immune cells, other is unknown. In neurodegenerative disease,
is there a hierarchy of responses? In a sea of complexity, the emergence of multiple immune-​related risk factor
it is often useful to build conceptual frameworks. The genes points to a potential cellular origin50,51, but the
evolving categorization of macrophage states provides a above examples argue strongly that the next frontier in
useful case study. The field previously adopted a bipolar glial–immune biology is to define the network activity
metric for microglial activation121, which rightly ignited of complex cellular responses and prioritize the most
great interest in the subject but was subsequently con- biologically important pathways.
sidered as an oversimplification215. It is being replaced by Though challenging, the complexity of immune cell–
the emerging concept of multidimensional state catego- glia interactions holds great opportunity for the develop-
rization, made increasingly possible by high throughput ment of targeted molecular therapies. Such efforts will
single-​cell analyses5–7,12,114,216–218. These methods have benefit greatly from new tools to extensively characterize
uncovered great complexity within canonically defined and interrogate complex cell–cell interactions in devel-
populations, but the reproducibility of ‘subpopulations’ opment and disease. These include cerebral organoids,
across samples, experiments and labs, in addition to their CRISPR screening, high content imaging and constantly
functional relevance, is still being established. improving sequencing technologies to map complex
With increased cellular resolution, our language pathways at single-​cell resolution. Recent glia–immune
for describing cell states has become more accurate cell-​specific advances include a 3D cell culture model
but, ironically, less concrete. At the level of microglial of Alzheimer disease220, the generation of microglia in
function, a similar pattern emerges. For example, in brain organoids221, massive single-​cell sequencing data-
the context of EAE, microglia-​derived TGFα reduces sets of microglia in health and disease6,7,216,222, and rap-
disease severity by signalling directly to astrocytes, idly developing analysis methods223. In the coming years,
whereas microglia-​derived VEGFβ increased disease these and other approaches stand to revolutionize our
burden90; microglia are thus both ‘good’ and ‘bad’, understanding of immune–glia interactions and how to
depending on what one prevents them for producing. modulate them.
Limiting astrocyte-​specific type I interferon signalling
in EAE increases microglial activation, suggesting that Published online xx xx xxxx

1. Barres, B. A. The mystery and magic of glia: myeloid subsets in health, aging, and disease. promote the formation of functional and structural
a perspective on their roles in health and disease. Immunity 48, 380–395 (2018). synapses.
Neuron 60, 430–440 (2008). This study provides an excellent description of the 26. Chuquet, J., Quilichini, P., Nimchinsky, E. A. &
2. Azevedo, F. A. C. et al. Equal numbers of neuronal different immune cell populations that reside Buzsáki, G. Predominant enhancement of glucose
and nonneuronal cells make the human brain an within the CNS using CyTOF. uptake in astrocytes versus neurons during activation
isometrically scaled-​up primate brain. J. Comp. 13. David, S., Kroner, A., Greenhalgh, A. D., Zarruk, J. G. of the somatosensory cortex. J. Neurosci. 30,
Neurol. 513, 532–541 (2009). & López-​Vales, R. Myeloid cell responses after spinal 15298–15303 (2010).
3. Morest, D. K. & Silver, J. Precursors of neurons, cord injury. J. Neuroimmunol. 321, 97–108 (2018). 27. Ioannou, M. S. et al. Neuron-​astrocyte metabolic
neuroglia, and ependymal cells in the CNS: 14. Herz, J., Filiano, A. J., Smith, A., Yogev, N. & Kipnis, J. coupling protects against activity-​induced fatty acid
what are they? Where are they from? How do Myeloid cells in the central nervous system. Immunity toxicity. Cell 177, 1522–1535 (2019).
they get where they are going? Glia 43, 6–18 46, 943–956 (2017). 28. Magistretti, P. J. & Allaman, I. Lactate in the brain:
(2003). 15. Filiano, A. J., Gadani, S. P. & Kipnis, J. How and why from metabolic end-​product to signalling molecule.
4. Saunders, A. et al. Molecular diversity and do T cells and their derived cytokines affect the injured Nat. Rev. Neurosci. 19, 235–249 (2018).
specializations among the cells of the adult mouse and healthy brain? Nat. Rev. Neurosci. 18, 375–384 29. Bergles, D. E. & Jahr, C. E. Synaptic activation of
brain. Cell 174, 1015–1030 (2018). (2017). glutamate transporters in hippocampal astrocytes.
5. Böttcher, C. et al. Human microglia regional 16. Kierdorf, K. & Prinz, M. Microglia in steady state. Neuron 19, 1297–1308 (1997).
heterogeneity and phenotypes determined by J. Clin. Invest. 127, 3201–3209 (2017). 30. Hertz, L. Possible role of neuroglia: a potassium-​
multiplexed single-​cell mass cytometry. Nat. Neurosci. 17. Pósfai, B., Cserép, C., Orsolits, B. & Dénes, Á. New mediated neuronal–neuroglial–neuronal impulse
22, 78–90 (2019). insights into microglia–neuron interactions: a neuron’s transmission system. Nature 206, 1091–1094
6. Masuda, T. et al. Spatial and temporal heterogeneity perspective. Neuroscience 405, 103–117 (2019). (1965).
of mouse and human microglia at single-​cell resolution. 18. Kisler, K., Nelson, A. R., Montagne, A. & Zlokovic, B. V. 31. Rusakov, D. A. Disentangling calcium-​driven
Nature 566, 388–392 (2019). Cerebral blood flow regulation and neurovascular astrocyte physiology. Nat. Rev. Neurosci. 16,
7. Van Hove, H. et al. A single-​cell atlas of mouse brain dysfunction in Alzheimer disease. Nat. Rev. Neurosci. 226–233 (2015).
macrophages reveals unique transcriptional identities 18, 419–434 (2017). 32. Abbott, N. J., Patabendige, A. A. K., Dolman, D. E. M.,
shaped by ontogeny and tissue environment. 19. Heindryckx, F. & Li, J.-P. Role of proteoglycans in Yusof, S. R. & Begley, D. J. Structure and function of
Nat. Neurosci. 22, 1021–1035 (2019). neuro-​inflammation and central nervous system the blood–brain barrier. Neurobiol. Dis. 37, 13–25
8. Grabert, K. et al. Microglial brain region-​dependent fibrosis. Matrix Biol. 68–69, 589–601 (2018). (2010).
diversity and selective regional sensitivities to aging. 20. Valny, M., Honsa, P., Kriska, J. & Anderova, M. 33. Santello, M., Toni, N. & Volterra, A. Astrocyte function
Nat. Neurosci. 19, 504–516 (2016). Multipotency and therapeutic potential of NG2 cells. from information processing to cognition and cognitive
This was the first study to show microglial Biochem. Pharmacol. 141, 42–55 (2017). impairment. Nat. Neurosci. 22, 154–166 (2019).
heterogeneity across brain regions. 21. Shechter, R., London, A. & Schwartz, M. Orchestrated 34. Hastings, M. H., Maywood, E. S. & Brancaccio, M.
9. Prinz, M., Erny, D. & Hagemeyer, N. Ontogeny and leukocyte recruitment to immune-​privileged sites: Generation of circadian rhythms in the suprachiasmatic
homeostasis of CNS myeloid cells. Nat. Immunol. 18, absolute barriers versus educational gates. Nat. Rev. nucleus. Nat. Rev. Neurosci. 19, 453–469 (2018).
385–392 (2017). Immunol. 13, 206–218 (2013). 35. Heneka, M. T., McManus, R. M. & Latz, E.
10. Kipnis, J. Multifaceted interactions between adaptive 22. Wohleb, E. S. & Godbout, J. P. Basic aspects of the Inflammasome signalling in brain function and
immunity and the central nervous system. Science immunology of neuroinflammation. Mod. Trends neurodegenerative disease. Nat. Rev. Neurosci. 19,
353, 766–771 (2016). Pharmacopsychiatry 28, 1–19 (2013). 610–621 (2018).
11. Goldmann, T. et al. Origin, fate and dynamics of 23. Zhou, B., Zuo, Y.-X. & Jiang, R.-T. Astrocyte morphology: 36. Anderson, M. A. et al. Astrocyte scar formation aids
macrophages at central nervous system interfaces. diversity, plasticity, and role in neurological diseases. central nervous system axon regeneration. Nature
Nat. Immunol. 17, 797–805 (2016). CNS Neurosci. Ther. 25, 665–673 (2019). 532, 195–200 (2016).
This study determined that non-​microglial 24. Allen, N. J. et al. Astrocyte glypicans 4 and 6 promote 37. Bradbury, E. J. et al. Chondroitinase ABC promotes
CNS macrophages arise from haematopoietic formation of excitatory synapses via GluA1 AMPA functional recovery after spinal cord injury. Nature
precursors during embryonic development receptors. Nature 486, 410–414 (2012). 416, 636–640 (2002).
and establish stable populations, with the 25. Christopherson, K. S. et al. Thrombospondins are 38. Anderson, M. A. et al. Required growth facilitators
notable exception of choroid plexus astrocyte-​secreted proteins that promote CNS propel axon regeneration across complete spinal cord
macrophages. synaptogenesis. Cell 120, 421–433 (2005). injury. Nature 561, 396–400 (2018).
12. Mrdjen, D. et al. High-​dimensional single-​cell mapping Together with reference 24, this study 39. Boulanger, J. J. & Messier, C. From precursors to
of central nervous system immune cells reveals distinct demonstrated that astrocyte-​secreted factors myelinating oligodendrocytes: contribution of intrinsic

Nature Reviews | Neuroscience


Reviews

and extrinsic factors to white matter plasticity in the 63. Radjavi, A., Smirnov, I., Derecki, N. & Kipnis, J. 87. Schwab, J. M., Zhang, Y., Kopp, M. A., Brommer, B.
adult brain. Neuroscience 269, 343–366 (2014). Dynamics of the meningeal CD4+ T-​cell repertoire are & Popovich, P. G. The paradox of chronic
40. Butt, A. M. in Encyclopedia of Neuroscience (ed. defined by the cervical lymph nodes and facilitate neuroinflammation, systemic immune suppression,
Squire, L. R.) 203–208 (Academic Press, 2009). cognitive task performance in mice. Mol. Psychiatry autoimmunity after traumatic chronic spinal cord
41. Zhu, X., Bergles, D. E. & Nishiyama, A. NG2 cells 19, 531–533 (2014). injury. Exp. Neurol. 258, 121–129 (2014).
generate both oligodendrocytes and gray matter 64. Quintana, E. et al. DNGR-1+ dendritic cells are located 88. Yun, S. P. et al. Block of A1 astrocyte conversion by
astrocytes. Development 135, 145–157 (2008). in meningeal membrane and choroid plexus of the microglia is neuroprotective in models of Parkinson’s
42. Dawson, M. R., Polito, A., Levine, J. M. & Reynolds, R. noninjured brain. Glia 63, 2231–2248 (2015). disease. Nat. Med. 24, 931–938 (2018).
NG2-expressing glial progenitor cells: an abundant 65. McCarthy, K. D. & de Vellis, J. Preparation of separate 89. Shinozaki, Y. et al. Transformation of astrocytes to
and widespread population of cycling cells in the adult astroglial and oligodendroglial cell cultures from rat a neuroprotective phenotype by microglia via P2Y1
rat CNS. Mol. Cell. Neurosci. 24, 476–488 (2003). cerebral tissue. J. Cell Biol. 85, 890–902 (1980). receptor downregulation. Cell Rep. 19, 1151–1164
43. Falcão, A. M. et al. Disease-​specific oligodendrocyte 66. Sievers, J., Parwaresch, R. & Wottge, H.-U. Blood (2017).
lineage cells arise in multiple sclerosis. Nat. Med. 24, monocytes and spleen macrophages differentiate 90. Rothhammer, V. et al. Microglial control of astrocytes
1837–1844 (2018). into microglia-​like cells on monolayers of astrocytes: in response to microbial metabolites. Nature 557,
44. Battefeld, A., Klooster, J. & Kole, M. H. P. Myelinating morphology. Glia 12, 245–258 (1994). 724–728 (2018).
satellite oligodendrocytes are integrated in a glial 67. Tanaka, J. & Maeda, N. Microglial ramification 91. Barres, B. A. et al. Cell death and control of cell
syncytium constraining neuronal high-​frequency requires nondiffusible factors derived from astrocytes. survival in the oligodendrocyte lineage. Cell 70,
activity. Nat. Commun. 7, 11298–11298 (2016). Exp. Neurol. 137, 367–375 (1996). 31–46 (1992).
45. Niu, J. et al. Aberrant oligodendroglial–vascular 68. Bohlen, C. J. et al. Diverse requirements for microglial 92. Nicholas, R. S. J., Wing, M. G. & Compston, A.
interactions disrupt the blood–brain barrier, survival, specification, and function revealed by defined- Nonactivated microglia promote oligodendrocyte
triggering CNS inflammation. Nat. Neurosci. 22, medium cultures. Neuron 94, 759–773.e8 (2017). precursor survival and maturation through the
709–718 (2019). 69. Foo, L. C. et al. Development of a method for the transcription factor NF-​κB. Eur. J. Neurosci. 13,
46. Greter, M., Lelios, I. & Croxford, A. L. Microglia versus purification and culture of rodent astrocytes. Neuron 959–967 (2001).
myeloid cell nomenclature during brain inflammation. 71, 799–811 (2011). 93. Hagemeyer, N. et al. Microglia contribute to normal
Front. Immunol. 6, 249–249 (2015). 70. Butovsky, O. et al. Identification of a unique TGF-​β myelinogenesis and to oligodendrocyte progenitor
47. Perry, V. H. & Gordon, S. Macrophages and microglia dependent molecular and functional signature in maintenance during adulthood. Acta Neuropathol.
in the nervous system. Trends Neurosci. 11, 273–277 microglia. Nat. Neurosci. 17, 131–143 (2014). 134, 441–458 (2017).
(1988). Together with references 110 and 111, this study 94. Schonberg, D. L. et al. Ferritin stimulates
48. Davalos, D. et al. ATP mediates rapid microglial showed that microglia can be defined by a specific oligodendrocyte genesis in the adult spinal cord and
response to local brain injury in vivo. Nat. Neurosci. 8, molecular signature in mice. can be transferred from macrophages to NG2 cells
752–758 (2005). 71. Greenhalgh, A. D. et al. Peripherally derived in vivo. J. Neurosci. 32, 5374–5384 (2012).
Together with reference 49, these studies were macrophages modulate microglial function to reduce 95. Wlodarczyk, A. et al. A novel microglial subset plays
the first to demonstrate the dynamic nature of inflammation after CNS injury. PLOS Biol. 16, a key role in myelinogenesis in developing brain.
microglial processes at steady state and in e2005264 (2018). EMBO J. 36, 3292–3308 (2017).
response to injury. 72. Gosselin, D. et al. An environment-​dependent 96. Liu, Y. et al. Concentration-​dependent effects of CSF1R
49. Nimmerjahn, A., Kirchhoff, F. & Helmchen, F. Resting transcriptional network specifies human microglia inhibitors on oligodendrocyte progenitor cells ex vivo
microglial cells are highly dynamic surveillants of brain identity. Science 356, eaal3222 (2017). and in vivo. Exp. Neurol. 318, 32–41 (2019).
parenchyma in vivo. Science 308, 1314–1318 (2005). This study comprehensively characterized the 97. Merrill, J. E., Ignarro, L. J., Sherman, M. P.,
50. Jansen, I. E. et al. Genome-​wide meta-​analysis transcriptional phenotype of human microglia. Melinek, J. & Lane, T. E. Microglial cell cytotoxicity
identifies new loci and functional pathways influencing 73. Cronk, J. C. et al. Peripherally derived macrophages of oligodendrocytes is mediated through nitric oxide.
Alzheimer’s disease risk. Nat. Genet. 51, 404–413 can engraft the brain independent of irradiation J. Immunol. 151, 2132–2141 (1993).
(2019). and maintain an identity distinct from microglia. 98. Zajicek, J., Wing, M., Scolding, N. & Compston, D.
51. Villegas-​Llerena, C., Phillips, A., Garcia-​Reitboeck, P., J. Exp. Med. 215, 1627–1647 (2018). Interactions between oligodendrocytes and microglia:
Hardy, J. & Pocock, J. M. Microglial genes regulating Together with reference 74, this study showed that a major role for complement and tumour necrosis
neuroinflammation in the progression of Alzheimer’s microglial transcriptomic identity is garnered by factor in oligodendrocyte adherence and killing.
disease. Curr. Opin. Neurobiol. 36, 74–81 (2016). both origin and environment. Brain 115, 1611–1631 (1992).
52. Schafer, D. P. et al. Microglia sculpt postnatal neural 74. Bennett, F. C. et al. A combination of ontogeny and 99. Moore, C. S. et al. Direct and indirect effects of
circuits in an activity and complement-​dependent CNS environment establishes microglial identity. immune and central nervous system-​resident cells on
manner. Neuron 74, 691–705 (2012). Neuron 98, 1170–1183 (2018). human oligodendrocyte progenitor cell differentiation.
Together with reference 53, this study demonstrated 75. Esmonde-​White, C. et al. Distinct function-​related J. Immunol. 194, 761–772 (2015).
the role of complement molecules in synapse molecular profile of adult human A2B5-positive 100. Miron, V. E. Microglia-​driven regulation of
elimination by microglia during development. pre-oligodendrocytes versus mature oligodendrocytes. oligodendrocyte lineage cells, myelination, and
53. Stevens, B. et al. The classical complement cascade J. Neuropathol. Exp. Neurol. 78, 468–479 (2019). remyelination. J. Leukoc. Biol. 101, 1103–1108
mediates CNS synapse elimination. Cell 131, 76. Healy, L. M., Yaqubi, M., Ludwin, S. & Antel, J. P. (2017).
1164–1178 (2007). Species differences in immune-​mediated CNS tissue 101. Miron, V. E. et al. M2 microglia and macrophages
54. Ginhoux, F. et al. Fate mapping analysis reveals that injury and repair: a (neuro)inflammatory topic. Glia drive oligodendrocyte differentiation during CNS
adult microglia derive from primitive macrophages. https://doi.org/10.1002/glia.23746(2019) (2019). remyelination. Nat. Neurosci. 16, 1211–1218 (2013).
Science 330, 841–845 (2010). 77. Même, W. et al. Proinflammatory cytokines released 102. Goldmann, T. et al. A new type of microglia gene
This study used elegant fate mapping approaches from microglia inhibit gap junctions in astrocytes: targeting shows TAK1 to be pivotal in CNS
to demonstrate the embryonic origin of microglia. potentiation by β-​amyloid. FASEB J. 20, 494–496 autoimmune inflammation. Nat. Neurosci. 16,
55. Schulz, C. et al. A lineage of myeloid cells independent (2006). 1618–1626 (2013).
of Myb and hematopoietic stem cells. Science 336, 78. Watanabe, M. et al. Th1 cells downregulate connexin 103. Wheeler, N. A. & Fuss, B. Extracellular cues influencing
86–90 (2012). 43 gap junctions in astrocytes via microglial activation. oligodendrocyte differentiation and (re)myelination.
56. Mass, E. et al. A somatic mutation in erythro-​myeloid Sci. Rep. 6, 38387 (2016). Exp. Neurol. 283, 512–530 (2016).
progenitors causes neurodegenerative disease. Nature 79. Retamal, M. A. et al. Cx43 hemichannels and gap 104. Giera, S. et al. The adhesion G protein-​coupled
549, 389–393 (2017). junction channels in astrocytes are regulated receptor GPR56 is a cell-​autonomous regulator of
This work showed that somatic mutations specifically oppositely by proinflammatory cytokines released oligodendrocyte development. Nat. Commun. 6, 6121
in embryonic progenitors of microglia can cause from activated microglia. J. Neurosci. 27, (2015).
neuronal loss, likely explaining why some patients 13781–13792 (2007). 105. Ackerman, S. D., Garcia, C., Piao, X., Gutmann, D. H.
with histiocytosis develop neurodegenerative 80. Kirkley, K. S., Popichak, K. A., Afzali, M. F., Legare, M. E. & Monk, K. R. The adhesion GPCR Gpr56 regulates
disease. & Tjalkens, R. B. Microglia amplify inflammatory oligodendrocyte development via interactions with
57. Kreutzberg, G. W. Microglia: a sensor for pathological activation of astrocytes in manganese neurotoxicity. Gα12/13 and RhoA. Nat. Commun. 6, 6122 (2015).
events in the CNS. Trends Neurosci. 19, 312–318 J. Neuroinflammation 14, 99 (2017). 106. Giera, S. et al. Microglial transglutaminase-2 drives
(1996). 81. Sofroniew, M. V. Astrogliosis. Cold Spring Harb. myelination and myelin repair via GPR56/ADGRG1
58. Beccari, S., Diaz-​Aparicio, I. & Sierra, A. Quantifying Perspect. Biol. 7, a020420 (2014). in oligodendrocyte precursor cells. eLife 7, e33385
microglial phagocytosis of apoptotic cells in the brain 82. Hwang, S.-Y. et al. Ionizing radiation induces astrocyte (2018).
in health and disease. Curr. Protoc. Immunol. 122, gliosis through microglia activation. Neurobiol. Dis. 107. Gibson, E. M. et al. Methotrexate chemotherapy
e49 (2018). 21, 457–467 (2006). induces persistent tri-​glial dysregulation that underlies
59. Korin, B. et al. High-​dimensional, single-​cell 83. Kyrkanides, S., Olschowka, J. A., Williams, J. P., chemotherapy-​related cognitive impairment. Cell 176,
characterization of the brain’s immune compartment. Hansen, J. T. & O’Banion, M. K. TNFα and IL-1β 43–55.e13 (2019).
Nat. Neurosci. 20, 1300–1309 (2017). mediate intercellular adhesion molecule-1 induction This work showed that glial cell dysfunction is the
60. Da Mesquita, S., Fu, Z. & Kipnis, J. The meningeal via microglia-​astrocyte interaction in CNS radiation likely cause of methotrexate-​induced cognitive
lymphatic system: a new player in neurophysiology. injury. J. Neuroimmunol. 95, 95–106 (1999). impairment.
Neuron 100, 375–388 (2018). 84. Liddelow, S. A. et al. Neurotoxic reactive astrocytes 108. Engelhardt, B., Vajkoczy, P. & Weller, R. O. The movers
61. Kivisäkk, P. et al. Human cerebrospinal fluid central are induced by activated microglia. Nature 541, and shapers in immune privilege of the CNS.
memory CD4+ T cells: evidence for trafficking through 481–487 (2017). Nat. Immunol. 18, 123–131 (2017).
choroid plexus and meninges via P-​selectin. Proc. Natl 85. Anderson, W. D. et al. Computational modeling of 109. Ransohoff, R. M. & Brown, M. A. Innate immunity
Acad. Sci. USA 100, 8389–8394 (2003). cytokine signaling in microglia. Mol. Biosyst. 11, in the central nervous system. J. Clin. Invest. 122,
62. Benakis, C., Llovera, G. & Liesz, A. The meningeal and 3332–3346 (2015). 1164–1171 (2012).
choroidal infiltration routes for leukocytes in stroke. 86. David, S. & Kroner, A. Repertoire of microglial and 110. Hickman, S. E. et al. The microglial sensome revealed
Ther. Adv. Neurol. Disord. 11, 1756286418783708 macrophage responses after spinal cord injury. by direct RNA sequencing. Nat. Neurosci. 16,
(2018). Nat. Rev. Neurosci. 12, 388–399 (2011). 1896–1905 (2013).

www.nature.com/nrn
Reviews

111. Bennett, M. L. et al. New tools for studying microglia 136. Shechter, R. et al. Recruitment of beneficial M2 159. Kurimoto, T. et al. Neutrophils express oncomodulin
in the mouse and human CNS. Proc. Natl Acad. Sci. macrophages to injured spinal cord is orchestrated by and promote optic nerve regeneration. J. Neurosci.
USA 113, E1738–E1746 (2016). remote brain choroid plexus. Immunity 38, 555–569 33, 14816–14824 (2013).
112. Ginhoux, F., Lim, S., Hoeffel, G., Low, D. & Huber, T. (2013). 160. Zenaro, E. et al. Neutrophils promote Alzheimer’s
Origin and differentiation of microglia. Front. Cell. 137. Kertser, A. et al. IFN-​γ-dependent activation of the disease-​like pathology and cognitive decline via LFA-1
Neurosci. 7, 45 (2013). brain’s choroid plexus for CNS immune surveillance integrin. Nat. Med. 21, 880–886 (2015).
113. Sousa, C. et al. Single-cell transcriptomics reveals and repair. Brain 136, 3427–3440 (2013). 161. Schwab, J. M., Chiang, N., Arita, M. & Serhan, C. N.
distinct inflammation-induced microglia signatures. 138. Kunis, G., Baruch, K., Miller, O. & Schwartz, M. Resolvin E1 and protectin D1 activate inflammation-​
EMBO Rep. 19, e46171 (2018). Immunization with a myelin-​derived antigen activates resolution programmes. Nature 447, 869–874
114. Krasemann, S. et al. The TREM2-APOE pathway the Brain’s choroid plexus for recruitment of (2007).
drives the transcriptional phenotype of dysfunctional immunoregulatory cells to the CNS and attenuates 162. Davies, C. L., Patir, A. & McColl, B. W. Myeloid cell
microglia in neurodegenerative diseases. Immunity disease progression in a mouse model of ALS. and transcriptome signatures associated with
47, 566–581 (2017). J. Neurosci. 35, 6381–6393 (2015). inflammation resolution in a model of self-​limiting
Together with references 216 and 217, this study 139. Szmydynger-​Chodobska, J. et al. Posttraumatic acute brain inflammation. Front. Immunol. 10, 1048
profiled microglia during development and injury, invasion of monocytes across the blood—cerebrospinal (2019).
and introduced the concept of a disease-​associated fluid barrier. J. Cereb. Blood Flow Metab. 32, 93–104 163. Serhan, C. N., Dalli, J., Colas, R. A., Winkler, J. W. &
microglial transcriptomic signature. (2012). Chiang, N. Protectins and maresins: New pro-​resolving
115. Butovsky, O. & Weiner, H. L. Microglial signatures and 140. Zarruk, J. G., Greenhalgh, A. D. & David, S. Microglia families of mediators in acute inflammation and
their role in health and disease. Nat. Rev. Neurosci. and macrophages differ in their inflammatory profile resolution bioactive metabolome. Biochim. Biophys.
19, 622–635 (2018). after permanent brain ischemia. Exp. Neurol. 301, Acta 1851, 397–413 (2015).
116. Shemer, A. et al. Engrafted parenchymal brain 120–132 (2018). 164. Francos-​Quijorna, I. et al. Maresin 1 promotes
macrophages differ from microglia in transcriptome, 141. Greenhalgh, A. D. & David, S. Differences in the inflammatory resolution, neuroprotection, and
chromatin landscape and response to challenge. phagocytic response of microglia and peripheral functional neurological recovery after spinal cord
Nat. Commun. 9, 5206 (2018). macrophages after spinal cord injury and its effects on injury. J. Neurosci. 37, 11731–11743 (2017).
117. Hines, D. J., Hines, R. M., Mulligan, S. J. & cell death. J. Neurosci. 34, 6316–6322 (2014). 165. Neumann, J. et al. Beware the intruder: real time
Macvicar, B. A. Microglia processes block the spread 142. Ajami, B., Bennett, J. L., Krieger, C., McNagny, K. M. observation of infiltrated neutrophils and neutrophil–
of damage in the brain and require functional chloride & Rossi, F. M. V. Infiltrating monocytes trigger EAE microglia interaction during stroke in vivo. PLOS ONE
channels. Glia 57, 1610–1618 (2009). progression, but do not contribute to the resident 13, e0193970 (2018).
118. Szalay, G. et al. Microglia protect against brain injury microglia pool. Nat. Neurosci. 14, 1142–1149 166. Gordon, S. Alternative activation of macrophages.
and their selective elimination dysregulates neuronal (2011). Nat. Rev. Immunol. 3, 23–35 (2003).
network activity after stroke. Nat. Commun. 7, 11499 143. Baruch, K. et al. PD-1 immune checkpoint blockade 167. Powell, D., Lou, M., Barros Becker, F. & Huttenlocher, A.
(2016). reduces pathology and improves memory in mouse Cxcr1 mediates recruitment of neutrophils and
119. Rice, R. A. et al. Elimination of microglia improves models of Alzheimer’s disease. Nat. Med. 22, supports proliferation of tumor-​initiating astrocytes
functional outcomes following extensive neuronal loss 135–137 (2016). in vivo. Sci. Rep. 8, 13285 (2018).
in the hippocampus. J. Neurosci. 35, 9977–9989 144. Prinz, M. & Priller, J. The role of peripheral 168. Hooshmand, M. J. et al. Neutrophils induce astroglial
(2015). immune cells in the CNS in steady state and disease. differentiation and migration of human neural stem
120. Bellver-​Landete, V. et al. Microglia are an essential Nat. Neurosci. 20, 136–144 (2017). cells via C1q and C3a synthesis. J. Immunol. 199,
component of the neuroprotective scar that forms 145. Guillot-​Sestier, M.-V. et al. IL10 deficiency rebalances 1069–1085 (2017).
after spinal cord injury. Nat. Commun. 10, 518 innate immunity to mitigate Alzheimer-​like pathology. 169. Ng, L. G., Ostuni, R. & Hidalgo, A. Heterogeneity of
(2019). Neuron 85, 534–548 (2015). neutrophils. Nat. Rev. Immunol. 19, 255–265 (2019).
121. Kigerl, K. A. et al. Identification of two distinct 146. Koronyo, Y. et al. Therapeutic effects of glatiramer 170. Filiano, A. J. et al. Unexpected role of interferon-​γ in
macrophage subsets with divergent effects causing acetate and grafted CD115+ monocytes in a mouse regulating neuronal connectivity and social behaviour.
either neurotoxicity or regeneration in the injured model of Alzheimer’s disease. Brain 138, 2399–2422 Nature 535, 425–429 (2016).
mouse spinal cord. J. Neurosci. 29, 13435–13444 (2015). 171. Derecki, N. C. et al. Regulation of learning and
(2009). 147. Wang, Y. et al. TREM2-mediated early microglial memory by meningeal immunity: a key role for IL-4.
122. Kroner, A. et al. TNF and increased intracellular iron response limits diffusion and toxicity of amyloid J. Exp. Med. 207, 1067–1080 (2010).
alter macrophage polarization to a detrimental M1 plaques. J. Exp. Med. 213, 667–675 (2016). 172. Murray, P. J. et al. Macrophage activation and
phenotype in the injured spinal cord. Neuron 83, 148. Fekete, R. et al. Microglia control the spread of polarization: nomenclature and experimental
1098–1116 (2014). neurotropic virus infection via P2Y12 signalling and guidelines. Immunity 41, 14–20 (2014).
123. Colonna, M. & Butovsky, O. Microglia function in recruit monocytes through P2Y12-independent 173. Filippi, M. et al. Multiple sclerosis. Nat. Rev.
the central nervous system during health and mechanisms. Acta Neuropathol. 136, 461–482 Dis. Primers 4, 43 (2018).
neurodegeneration. Annu. Rev. Immunol. 35, (2018). 174. Dobson, R. & Giovannoni, G. Multiple sclerosis –
441–468 (2017). 149. Grønberg, N. V., Johansen, F. F., Kristiansen, U. & a review. Eur. J. Neurol. 26, 27–40 (2019).
124. Rock, R. B. et al. Role of microglia in central nervous Hasseldam, H. Leukocyte infiltration in experimental 175. Li, R., Patterson, K. R. & Bar-​Or, A. Reassessing B cell
system infections. Clin. Microbiol. Rev. 17, 942–964 stroke. J. Neuroinflammation 10, 115 (2013). contributions in multiple sclerosis. Nat. Immunol. 19,
(2004). 150. Taylor, R. A. et al. TGF-​β1 modulates microglial 696–707 (2018).
125. Neher, J. J. & Cunningham, C. Priming microglia for phenotype and promotes recovery after intracerebral 176. Brennan, F. H. & Popovich, P. G. Emerging targets
innate immune memory in the brain. Trends Immunol. hemorrhage. J. Clin. Invest. 127, 280–292 (2017). for reprograming the immune response to promote
40, 358–374 (2019). 151. Yamasaki, R. et al. Differential roles of microglia and repair and recovery of function after spinal cord injury.
126. Lee, H.-M., Kang, J., Lee, S. J. & Jo, E.-K. Microglial monocytes in the inflamed central nervous system. Curr. Opin. Neurol. 31, 334–344 (2018).
activation of the NLRP3 inflammasome by the priming J. Exp. Med. 211, 1533–1549 (2014). 177. Raposo, C. et al. CNS repair requires both effector and
signals derived from macrophages infected with 152. Anderson, W. D., Greenhalgh, A. D., Takwale, A., regulatory T cells with distinct temporal and spatial
mycobacteria. Glia 61, 441–452 (2013). David, S. & Vadigepalli, R. Novel influences of IL-10 profiles. J. Neurosci. 34, 10141–10155 (2014).
127. Qin, Y. et al. Macrophage-​microglia networks drive M1 on CNS inflammation revealed by integrated analyses 178. Vindegaard, N. et al. T-​cells and macrophages peak
microglia polarization after mycobacterium infection. of cytokine networks and microglial morphology. weeks after experimental stroke: spatial and temporal
Inflammation 38, 1609–1616 (2015). Front. Cell. Neurosci. 11, 233 (2017). characteristics. Neuropathology 37, 407–414 (2017).
128. Renner, N. A. et al. Microglia activation by SIV-​infected 153. Sharma, S. et al. Bone marrow mononuclear cells 179. Cramer, J. V., Benakis, C. & Liesz, A. T cells in the
macrophages: alterations in morphology and cytokine protect neurons and modulate microglia in cell culture post-​ischemic brain: troopers or paramedics?
secretion. J. Neurovirol. 18, 213–221 (2012). models of ischemic stroke. J. Neurosci. Res. 88, J. Neuroimmunol. 326, 33–37 (2019).
129. Wolfe, H., Minogue, A. M., Rooney, S. & Lynch, M. A. 2869–2876 (2010). 180. Zrzavy, T. et al. Dominant role of microglial and
Infiltrating macrophages contribute to age-​related 154. Shechter, R. et al. Infiltrating blood-​derived macrophage innate immune responses in human
neuroinflammation in C57/BL6 mice. Mechanisms macrophages are vital cells playing an anti-​ ischemic infarcts. Brain Pathol. 28, 791–805 (2018).
Ageing Dev. 173, 84–91 (2018). inflammatory role in recovery from spinal cord injury 181. Fleming, J. C. et al. The cellular inflammatory
130. Shechter, R. & Schwartz, M. Harnessing monocyte-​ in mice. PLOS Med. 6, e1000113 (2009). response in human spinal cords after injury.
derived macrophages to control central nervous 155. Frik, J. et al. Cross-talk between monocyte invasion Brain 129, 3249–3269 (2006).
system pathologies: no longer ‘if’ but ‘how’. J. Pathol. and astrocyte proliferation regulates scarring in brain 182. Ito, M. et al. Brain regulatory T cells suppress
229, 332–346 (2013). injury. EMBO Rep. 19, e45294 (2018). astrogliosis and potentiate neurological recovery.
131. David, S., López-​Vales, R. & Wee Yong, V. in Handbook 156. Haan, N., Zhu, B., Wang, J., Wei, X. & Song, B. Nature 565, 246–250 (2019).
of Clinical Neurology Vol. 109 (eds Verhaagen, J. Crosstalk between macrophages and astrocytes 183. Unger, M. S. et al. Doublecortin expression in CD8+
& McDonald, J. W.) 485–502 (Elsevier, 2012). affects proliferation, reactive phenotype and T-cells and microglia at sites of amyloid-​β plaques:
132. Kim, E. & Cho, S. Microglia and monocyte-​derived inflammatory response, suggesting a role during a potential role in shaping plaque pathology?
macrophages in stroke. Neurotherapeutics 13, reactive gliosis following spinal cord injury. Alzheimers Dement. 14, 1022–1037 (2018).
702–718 (2016). J. Neuroinflammation 12, 109 (2015). 184. Merlini, M., Kirabali, T., Kulic, L., Nitsch, R. M. &
133. Hu, X. et al. Microglial and macrophage polarization 157. Andjelkovic, A. V., Kerkovich, D. & Pachter, J. S. Ferretti, M. T. Extravascular CD3+ T cells in brains of
— new prospects for brain repair. Nat. Rev. Neurol. Monocyte:astrocyte interactions regulate MCP-1 Alzheimer disease patients correlate with tau but not
11, 56–64 (2015). expression in both cell types. J. Leukoc. Biol. 68, with amyloid pathology: an immunohistochemical
134. McKee, C. A. & Lukens, J. R. Emerging roles for 545–552 (2000). study. Neurodegener. Dis. 18, 49–56 (2018).
the immune system in traumatic brain injury. 158. Harris, J. E. et al. Monocyte-​astrocyte networks 185. Togo, T. et al. Occurrence of T cells in the brain of
Front. Immunol. 7, 556–556 (2016). regulate matrix metalloproteinase gene expression Alzheimer’s disease and other neurological diseases.
135. Ge, R. et al. Choroid plexus-​cerebrospinal fluid route and secretion in central nervous system tuberculosis J. Neuroimmunol. 124, 83–92 (2002).
for monocyte-​derived macrophages after stroke. in vitro and in vivo. J. Immunol. 178, 1199–1207 186. Brochard, V. et al. Infiltration of CD4+ lymphocytes
J. Neuroinflammation 14, 153 (2017). (2007). into the brain contributes to neurodegeneration in a

Nature Reviews | Neuroscience


Reviews

mouse model of Parkinson disease. J. Clin. Invest. 207. McGeer, P. L. & McGeer, E. G. Inflammatory processes 230. Haimon, Z. et al. Re-evaluating microglia expression
119, 182–192 (2009). in amyotrophic lateral sclerosis. Muscle Nerve 26, profiles using RiboTag and cell isolation strategies.
187. Troost, D., van den Oord, J. J. & Vianney de Jong, J. M. 459–470 (2002). Nat. Immunol. 19, 636–644 (2018).
Immunohistochemical characterization of the 208. Chiu, I. M. et al. T lymphocytes potentiate endogenous 231. van den Brink, S. C. et al. Single-cell sequencing
inflammatory infiltrate in amyotrophic lateral sclerosis. neuroprotective inflammation in a mouse model of reveals dissociation-​induced gene expression in
Neuropathol. Appl. Neurobiol. 16, 401–410 (1990). ALS. Proc. Natl Acad. Sci. USA 105, 17913–17918 tissue subpopulations. Nat. Methods 14, 935–936
188. Engelhardt, J. I., Tajti, J. & Appel, S. H. Lymphocytic (2008). (2017).
infiltrates in the spinal cord in amyotrophic lateral 209. Sheean, R. K. et al. Association of regulatory T-​cell 232. Bonnardel, J. et al. Stellate cells, hepatocytes, and
sclerosis. Arch. Neurol. 50, 30–36 (1993). expansion with progression of amyotrophic lateral endothelial cells imprint the Kupffer cell identity on
189. Beers, D. R. & Appel, S. H. Immune dysregulation sclerosis: a study of humans and a transgenic mouse monocytes colonizing the liver macrophage niche.
in amyotrophic lateral sclerosis: mechanisms and model. JAMA Neurol. 75, 681–689 (2018). Immunity 51, 638–654 (2019).
emerging therapies. Lancet Neurol. 18, 211–220 210. Korhonen, P. et al. Long-​term interleukin-33 treatment 233. Srinivasan, R. et al. New transgenic mouse lines for
(2019). delays disease onset and alleviates astrocytic selectively targeting astrocytes and studying calcium
190. González, H., Contreras, F. & Pacheco, R. Regulation of activation in a transgenic mouse model of amyotrophic signals in astrocyte processes in situ and in vivo.
the neurodegenerative process associated to Parkinson’s lateral sclerosis. IBRO Rep. 6, 74–86 (2019). Neuron 92, 1181–1195 (2016).
disease by CD4+ T-​cells. J. Neuroimmune Pharmacol. 211. Gadani, S. P. et al. The glia-​derived alarmin IL-33 234. Sun, L. O. et al. Spatiotemporal control of CNS
10, 561–575 (2015). orchestrates the immune response and promotes myelination by oligodendrocyte programmed cell
191. van Dyck, C. H. Anti-​amyloid-β monoclonal antibodies recovery following CNS injury. Neuron 85, 703–709 death through the TFEB-​PUMA axis. Cell 175,
for Alzheimer’s disease: pitfalls and promise. Biol. (2015). 1811–1826 (2018).
Psychiatry 83, 311–319 (2018). 212. Harms, A. S. et al. MHCII is required for α-​synuclein- 235. von Jonquieres, G. et al. Glial promoter selectivity
192. Bachmann, M. F., Jennings, G. T. & Vogel, M. A vaccine induced activation of microglia, CD4 T cell proliferation, following AAV-​delivery to the immature brain. PLOS
against Alzheimer’s disease: anything left but faith? and dopaminergic neurodegeneration. J. Neurosci. 33, ONE 8, e65646 (2013).
Expert Opin. Biol. Ther. 19, 73–78 (2019). 9592–9600 (2013). 236. Varvel, N. H. et al. Microglial repopulation model
193. Bodhankar, S., Chen, Y., Vandenbark, A. A., 213. Sommer, A. et al. Infiltrating T lymphocytes reduce reveals a robust homeostatic process for replacing
Murphy, S. J. & Offner, H. Treatment of experimental myeloid phagocytosis activity in synucleinopathy CNS myeloid cells. Proc. Natl Acad. Sci. USA 109,
stroke with IL-10-producing B-​cells reduces infarct size model. J. Neuroinflammation 13, 174 (2016). 18150–18155 (2012).
and peripheral and CNS inflammation in wild-​type 214. Fan, K.-Q. et al. Stress-​induced metabolic disorder in 237. Heppner, F. L. et al. Experimental autoimmune
B-cell-sufficient mice. Metab. Brain Dis. 29, 59–73 peripheral CD4+ T cells leads to anxiety-​like behavior. encephalomyelitis repressed by microglial paralysis.
(2014). Cell 179, 864–879 (2019). Nat. Med. 11, 146–152 (2005).
194. Ankeny, D. P., Guan, Z. & Popovich, P. G. B cells This study showed that T cell-​mediated actions on 238. Kitic, M., See, P., Bruttger, J., Ginhoux, F. &
produce pathogenic antibodies and impair recovery oligodendrocytes in a specific brain region affect Waisman, A. in Microglia: Methods and Protocols
after spinal cord injury in mice. J. Clin. Invest. 119, behaviour. (eds Garaschuk, O. & Verkhratsky, A.) 217–230
2990–2999 (2009). 215. Ransohoff, R. M. A polarizing question: do M1 and (Springer, 2019).
195. Doyle, K. P. et al. B-​lymphocyte-mediated delayed M2 microglia exist? Nat. Neurosci. 19, 987–991 239. Elmore, M. R. P. et al. Colony-​stimulating factor 1
cognitive impairment following stroke. J. Neurosci. 35, (2016). receptor signaling is necessary for microglia viability,
2133–2145 (2015). 216. Hammond, T. R. et al. Single-​cell RNA sequencing of unmasking a microglia progenitor cell in the adult
196. Westendorp, W. F., Nederkoorn, P. J., Vermeij, J. D., microglia throughout the mouse lifespan and in the brain. Neuron 82, 380–397 (2014).
Dijkgraaf, M. G. & de Beek, D. V. Post-​stroke infection: injured brain reveals complex cell-​state changes. This was the first study to use a CSF1R antagonist
a systematic review and meta-​analysis. BMC Neurol. Immunity 50, 253–271 (2019). to ablate microglia from the adult mouse brain.
11, 110 (2011). 217. Keren-​Shaul, H. et al. A unique microglia type 240. Han, J., Harris, R. A. & Zhang, X.-M. An updated
197. Sun, G. et al. γδ T cells provide the early source associated with restricting development of Alzheimer’s assessment of microglia depletion: current concepts
of IFN-γ to aggravate lesions in spinal cord injury. disease. Cell 169, 1276–1290 (2017). and future directions. Mol. Brain 10, 25 (2017).
J. Exp. Med. 215, 521–535 (2018). 218. O’Koren, E. G. et al. Microglial function is distinct 241. Rojo, R. et al. Deletion of a Csf1r enhancer selectively
198. Liesz, A. et al. Regulatory T cells are key in different anatomical locations during retinal impacts CSF1R expression and development of tissue
cerebroprotective immunomodulators in acute homeostasis and degeneration. Immunity 50, macrophage populations. Nat. Commun. 10, 3215
experimental stroke. Nat. Med. 15, 192–199 (2009). 723–737.e7 (2019). (2019).
This was one of the first studies to show a 219. Rothhammer, V. et al. Type I interferons and microbial 242. The Jackson Laboratory. Immunodeficient mouse and
protective role of Treg cells in experimental metabolites of tryptophan modulate astrocyte activity xenograft host comparisons. Jackson Lab. https://
stroke. and central nervous system inflammation via the aryl www.jax.org/jax-​mice-and-services/find-and-order-jax-
199. Na, S.-Y., Mracsko, E., Liesz, A., Hünig, T. & hydrocarbon receptor. Nat. Med. 22, 586–597 (2016). mice/most-​popular-jax-​mice-strains/immunodeficient-​
Veltkamp, R. Amplification of regulatory T cells using 220. Park, J. et al. A 3D human triculture system modeling mouse-and-​xenograft-host-​comparisons (2019).
a CD28 superagonist reduces brain damage after neurodegeneration and neuroinflammation in 243. Herz, J. et al. Role of neutrophils in exacerbation
ischemic stroke in mice. Stroke 46, 212–220 Alzheimer’s disease. Nat. Neurosci. 21, 941–951 of brain injury after focal cerebral ischemia in
(2015). (2018). hyperlipidemic mice. Stroke 46, 2916–2925
200. Xie, L. et al. mTOR signaling inhibition modulates 221. Ormel, P. R. et al. Microglia innately develop within (2015).
macrophage/microglia-​mediated neuroinflammation cerebral organoids. Nat. Commun. 9, 4167 (2018).
and secondary injury via regulatory T cells after focal 222. Li, Q. et al. Developmental heterogeneity of microglia Acknowledgements
ischemia. J. Immunol. 192, 6009–6019 (2014). and brain myeloid cells revealed by deep single-​cell The authors thank the all the scientists involved in producing
201. Xie, L., Choudhury, G. R., Winters, A., Yang, S. H. & RNA sequencing. Neuron 101, 207–223 (2019). the original work and we hope we have done justice to their
Jin, K. Cerebral regulatory T cells restrain microglia/ 223. Cohen, M. et al. Lung single-​cell signaling interaction findings.
macrophage-​mediated inflammatory responses via map reveals basophil role in macrophage imprinting.
IL-10. Eur. J. Immunol 45, 180–191 (2015). Cell 175, 1031–1044 (2018).
Author contributions
202. Wu, B. et al. Improved regeneration after spinal cord 224. Guttenplan, K. A. & Liddelow, S. A. Astrocytes and
A.D.G. and F.C.B. researched data for the article and made
injury in mice lacking functional T- and B-​lymphocytes. microglia: models and tools. J. Exp. Med. 216, 71–83
substantial contributions to the discussion of the content. All
Exp. Neurol. 237, 274–285 (2012). (2019).
the authors contributed equally to writing the article and to the
203. Späni, C. et al. Reduced β-​amyloid pathology in an 225. Bohlen, C. J., Bennett, F. C. & Bennett, M. L. Isolation
review and editing of the manuscript before submission.
APP transgenic mouse model of Alzheimer’s disease and culture of microglia. Curr. Protoc. Immunol. 125,
lacking functional B and T cells. Acta Neuropathol. e70 (2019).
Commun. 3, 71 (2015). 226. Qian, X., Song, H. & Ming, G. L. Brain organoids: Competing interests
204. Browne, T. C. et al. IFN-​γ production by amyloid advances, applications and challenges. Development The authors declare no competing interests.
β-specific Th1 cells promotes microglial activation 146, dev166074 (2019).
and increases plaque burden in a mouse model of 227. Marton, R. M. et al. Differentiation and maturation of Peer review information
Alzheimer’s disease. J. Immunol. 190, 2241–2251 oligodendrocytes in human three-​dimensional neural Nature Reviews Neuroscience thanks O. Butovsky, J. Schwab
(­20­13­). cultures. Nat. Neurosci. 22, 484–491 (2019). and B. Stevens for their contribution to the peer review of this
205. Marsh, S. E. et al. The adaptive immune system 228. Zhang, Y. et al. An RNA-​sequencing transcriptome and work.
restrains Alzheimer’s disease pathogenesis by splicing database of glia, neurons, and vascular cells of
modulating microglial function. Proc. Natl Acad. the cerebral cortex. J. Neurosci. 34, 11929–11947 Publisher’s note
Sci. USA 113, E1316–E1325 (2016). (2014). Springer Nature remains neutral with regard to jurisdictional
206. Pasinelli, P. & Brown, R. H. Molecular biology of 229. Buttgereit, A. et al. Sall1 is a transcriptional regulator claims in published maps and institutional affiliations.
amyotrophic lateral sclerosis: insights from genetics. defining microglia identity and function. Nat. Immunol.
Nat. Rev. Neurosci. 7, 710–723 (2006). 17, 1397–1406 (2016). © Springer Nature Limited 2020

www.nature.com/nrn

You might also like