Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Aug 1.
Published in final edited form as: Trends Cell Biol. 2016 Mar 20;26(8):587–597. doi: 10.1016/j.tcb.2016.02.006

Microglia: Architects of the Developing Nervous System

Jeffrey L Frost 1, Dorothy P Schafer 1,*
PMCID: PMC4961529  NIHMSID: NIHMS763988  PMID: 27004698

Abstract

Microglia are resident macrophages of the central nervous system (CNS), representing 5–10% of total CNS cells. Recent findings reveal that microglia enter the embryonic brain, take up residence before the differentiation of other CNS cell types, and become critical regulators of CNS development. Here, we discuss exciting new work implicating microglia in a range of developmental processes including regulation of cell number and spatial patterning of CNS cells, myelination, and formation and refinement of neural circuits. Furthermore, we review studies suggesting that these cellular functions may result in modulation of behavior, which has important implications for a variety of neurological disorders.

Keywords: Microglia, Development, Central Nervous System

Microglia in the developing central nervous system

Development represents a remarkably dynamic window in the course of an organism’s life, requiring coordination and communication among vastly different organ systems and cell types. In the central nervous system (CNS), a large variety of neurons and glial cells must communicate with one another to achieve the exquisite structure and function that are characteristic of the mature system. Included among these cell types are microglia, the resident brain macrophage, which make up approximately 5–10% of total CNS cells. Microglia are one of the first tissue macrophages to be born in the yolk sac at ~embryonic day 7.5 (E7.5) and migrate into the brain rudiment at ~E9.5 where they take up residence and self-renew throughout life [15]. The timing of this colonization occurs prior to the differentiation of other resident nervous system cells [6]. As a result, microglia are present at the right time and place to play critical roles in CNS development. Here, we review recent work demonstrating that microglia regulate an array of developmental processes that are necessary for achieving appropriate cellular architecture and function in the mature CNS. We also discuss emerging idea that these cellular functions may be novel mechanisms by which devastating neurological disorders manifest (Table 1).

Table 1.

Animal models or human diseases associated with microglia dysfunction during development.

Model or Disease Species Manipulation Behavior Affected Ref
Animal Studies
Early life infection followed by later life immune challenge Rat Prenatal Escherichia coli and later life LPS challenge +/− handling Memory [7274]
Maternal immune activation Mouse or nonhuman primate Poly(I:C) in utero +/− postnatal stress Anxiety, sensorimotor gating, learning, psychotomimetic drug sensitivity, social interactions, ultrasonic vocalizations, repetitive behavior [7578]
Estradiol-treated females Rat Minocycline Copulatory behavior [79]
CX3CR1 KO Mouse Gene ablation Social interactions [80]
Microglia depletion in juvenile Mouse Diphtheria toxin receptor expression using CX3CR1cre-ERT2 Motor learning [85]
Microglia-specific BDNF null Mouse Gene ablation using CX3CR1cre-ERT2 Motor learning [85]
Hoxb8 mutant Mouse Gene mutation +/− wild-type myeloid cells Grooming [81]
Mecp2 null Mouse Gene mutation +/− wild-type myeloid cells Locomotion, weight, breathing, and lifespan [82, 83]
Mecp2 null Mouse Gene mutation +/− wild-type myeloid cells None [84]
Human Studies
HDLS Human CSF1R mutations Impairments in mood, social interactions, cognition, and motor control [95]
Nasu-Hakola disease Human DAP12 or TREM2 mutations Psychosis and dementia [96]
Frontotemporal Dementia Human TREM2 mutation Dementia [100]
Alzheimer’s disease Human CD33 risk allele; TREM2 variant Increased disease susceptibility [9799]
Multiple Sclerosis Human IRF8 and TNFRSF1A variants Increased disease susceptibility [101, 102]
Bipolar and Major Depression Human P2RX7 risk allele Increased disease susceptibility [103, 104]
Schizophrenia human C4 variant Increased disease susceptibility [105]

Control of Neuron Cell Fate and Number

In the developing CNS, resident cells are born and migrate to their appropriate location. During this process, a subset of newly born cells must be lost during normal programmed cell death (NPCD) while the remaining cells mature [7, 8]. Early imaging studies demonstrated that microglia engulf dead or dying cells throughout the developing brain [9, 10] (Figure 1A). However, it remained unclear whether microglia played a more active role by initiating the cell death program prior to engulfment. Some of the most direct evidence for a more active role were in vitro studies in chick retina where NPCD was reduced when retinas were cultured in the absence of microglia [11]. When microglia were added back to the cultures, retinal cell death increased—an effect attributed to microglia-derived nerve growth factor (NGF). Similarly, in cultured mouse cerebellar slices or rat spinal cord explants, microglia engulfed dead or dying cells, and pharmacological depletion of microglia resulted in reduced Purkinje neuron and motoneuron NPCD [12, 13]. Superoxide ions released from microglia mediated Purkinje neuron cell death in the cerebellum, while microglial-derived TNF-α initiated NPCD of motoneurons in the spinal cord. These data suggest that microglia not only play a critical role in clearing cellular debris of dead or dying cells, but also actively initiate the cell death program.

Figure 1. Microglia interact with developing cells in the postnatal brain.

Figure 1

A, Microglia (green) in the juvenile (P30) mouse hippocampus represent 5–10% of total CNS cells. Microglia are labeled using a transgenic reporter (CX3CR1egfp/WT) and neurons are labeled with an antibody directed against NeuN (purple). Scale bar=100µm. B, Microglia (Cx3CR1egfp/WT, green) in the SVZ of a P13 mouse engulfing actively dividing cells labeled with 5-ethynyl-2’-deoxyuridine (EDU; purple). Often these apoptotic, dividing cells are found enveloped within microglial processes that form phagocytic cups (arrow and enlarged in inset). C, Microglia (CX3CR1egfp/WT, green) closely associate and often contact (arrow and inset) retinal ganglion cell (RGC) presynaptic inputs labeled by anterograde tracing with cholera toxin β subunit conjugated to Alexa 594 (CTB-594, purple) in the juvenile mouse lateral geniculate nucleus (LGN, P29). Di, Microglia (CX3CR1egfp/WT, green) in the early postnatal LGN (P5) closely associate with RGC presynaptic inputs (CTB-594, purple). Dii, Engulfment of presynaptic inputs can be visualized within the microglia soma and processes (arrow, inset) once all RGC input fluorescence outside the microglia volume is subtracted. B–D, Scale bar=10µm.

Similar to in vitro studies, microglia have been suggested to regulate NPCD at sites of neurogenesis in vivo. In developing zebrafish, phosphatidyl serine receptors, Bal1 and Tim-4, were recently identified to regulate the phagocytic machinery necessary for microglia to clear dying neurons in the developing brain [14]. However, apoptosis still progressed in the absence of microglia. In macaque monkeys and rats, microglia engulfed excess neural progenitor cells (NPCs) as neurogenesis neared completion in the cerebral cortex [15]. Furthermore, the number of cortical NPCs increased when microglia were pharmacologically inactivated with broad spectrum antibiotics (minocycline or doxycycline) or depleted with liposomal clodronate [15]. Conversely, treating mice in utero with lipopolysaccharide (LPS) to increase the inflammatory state of microglia resulted in a decrease in NPCs in the cortex. These data suggest that microglia regulate NPC number by initiating cell death in the mammalian brain and engulfing dead or dying cells. Given that the pharmacological approaches in the mammal are relatively non-specific, future work is necessary to determine if these effects are microglia specific. In addition, it is unknown whether Bal1 and Tim-4-dependent phagocytosis of dying neurons in the developing zebrafish is a conserved mechanism across species.

Microglial-derived factors are also critical to survival, proliferation, and maturation of NPCs in the developing brain. For example, the addition of microglia-conditioned media to cultured neurons resulted in an increase in NPC proliferation coupled with enhanced neuron survival and maturation [1618]. Similarly, NPCs isolated from embryonic day 12 (E12) mice that lack microglia (PU.1-deficient mice) exhibited decreases in both proliferation and astrogenesis, effects that were attenuated by the addition of wild-type microglia [19]. In contrast, NPCs isolated from 3 month-old rats and co-cultured with increasing concentrations of microglia, revealed an inverse correlation between progenitor cell survival and microglia concentration [20]. These different conclusions may be a result of regional differences in microglia function (cortex vs. hippocampus) or differences in culture preparation. To assess whether microglia provide trophic support to neurons in vivo, genetic mouse models have been utilized. Mice deficient in the fractalkine receptor (CX3CR1), a chemokine receptor highly enriched in microglia in healthy CNS, had significant increases in numbers of apoptotic neurons in layer V of the postnatal cerebral cortex [21]. This effect was replicated by pharmacologically inactivating or genetically depleting microglia. Furthermore, because similar rates of apoptotic cell engulfment were observed in CX3CR1-deficient and wild-type microglia, it is unlikely that increased apoptotic neurons resulted from inefficient clearance of dead cells by CX3CR1-deficient microglia. Instead, this effect was attributed to reductions in insulin-like growth factor 1 (IGF-1) signaling, a potent trophic factor for NPC survival [2224], in CX3CR1-deficient mice. In another study, pharmacological inactivation of microglia with minocycline in postnatal rats caused a reduction in the numbers of proliferating progenitor cells and mature oligodendrocytes in the subventircular zone (SVZ). [25] In vitro experiments on cultured neurospheres suggest that this effect is regulated by microglia-derived cytokines including interleukin (IL)-1β, interferon-γ, and IL-6.

While microglia regulate neuronal cell number throughout the brain by initiating NPCD and engulfing dead or dying cells, other work has demonstrated a concomitant function to provide trophic support to progenitor cells (Figure 2, Key Figure). How does a cell simultaneously promote cell death, proliferation, survival, and maturation? Do these functions represent regional differences in neuronal receptivity or heterogeneity of microglia in the brain? Answers to these questions will be important to identify the function of microglia in the developing brain in health and disease.

Figure 2. Key Figure. A summary of microglia functions in the developing brain.

Figure 2

New data demonstrate that microglia can affect the development of other resident CNS cell types throughout the CNS. NPC, neural precursor cell; OPC, oligodendrocyte precursor cell.

Regulation of non-neuronal cell development

In addition to neurons, microglia have been implicated in the development of other resident CNS cell types (i.e., other glial cells). For example, in vitro evidence suggests that microglia-conditioned media increases the differentiation of neural stem/precursor cells (NSPCs) into astrocytes through IL-6 and leukaemia inhibitory factor (LIF) [26]. Similarly, microglia-conditioned media promotes survival and differentiation of cultured oligodendrocyte precursor cells (OPCs) into mature oligodendrocytes through a number of secreted factors including IGF-1, nuclear factor kappaB (NF-κB), IL-1β, and IL-6 [25, 2729]. Microglia have also been suggested to promote myelination by providing iron, a necessary co-factor for myelination, to oligodendrocytes [3032]. These results suggest that microglia have the potential to regulate survival, proliferation, and maturation of most developing CNS cell types. However, the majority of work assessing the role of microglia in non-neuronal cell development is in vitro. It is unknown whether these mechanisms apply in vivo.

There is also mounting evidence that microglia regulate vascularization of the nervous system. Initial observations showed that microglia arrive in the CNS before blood vessels develop and closely associate with invading vessels, particularly in the developing retina [33, 34]. In vitro and in vivo depletion of microglia in rodents has further suggested that microglia are necessary for vascular branching in the retina and hindbrain [3538]. However, other studies have reported the opposite effect--depletion of microglia in an ex vivo retinal preparation or in vivo loss of microglial Wnt-ligand transporter (Wntless) expression results in increased vascular branching [39, 40]. Thus, future work is necessary to determine precisely how microglia regulate vasculogenesis. In addition, it remains unknown if microglia-dependent regulation of vascular branching is restricted to the developing retina or whether this is a broader effect occurring throughout the CNS.

Activity-dependent patterning and maturation of neural circuits

Neurons initially connect with each other at synapses to form a crude wiring diagram. Neural activity then regulates the remodeling and maturation of this immature synaptic connectivity whereby less active synaptic connections are eliminated and more active connections are maintained and strengthened [41, 42]. Microglia express neurotransmitter receptors and live imaging studies have revealed that these cells are dynamic sensors of neural activity [4346]. For example, activity-dependent release of ATP from neurons regulates microglial process motility and outgrowth [4750] and dampening neural activity int he visual cortex by rearing mice in the dark results in decreased microglial process motility [51]. Furthermore, increasing or decreasing activity in the visual cortex changes the frequency and duration of microglial contact with synapses and induces engulfment of elements that resemble synapses by ultrastructure [51, 52]. These data suggest that microglia could regulate synapse development through activity-dependent mechanisms.

To understand the functional consequences of activity-dependent microglial responses and physical interactions with synapses (Figure 1B–C), recent data has suggested key roles for microglia in regulating maturation and remodeling of synaptic connectivity. Earlier work in acute hippocampal slices prepared from mice mutant for the microglial transmembrane receptor, DNAX-activation protein 12 (DAP12), demonstrated an increase in electrophysiological features characteristic of less mature synapses [53]. In a follow-up study, these DAP12 mutant mice also displayed abnormalities in the development of structural synapses in the hippocampus [54]. In more recent studies, transient reductions in microglia numbers in the hippocampus or barrel cortex due to a genetic deletion of CX3CR1 (CX3CR1 KO) resulted in delayed maturation of structural and functional synapses [55, 56]. These effects are later attenuated in juvenile CX3CR1 KO mice after microglia density reaches wild-type levels. Together, these data suggest that microglia regulate the maturation of synapses in the postnatal brain.

Microglia have also been implicated in the remodeling of developing synapses in response to changes in neural activity. Using the developing mouse retinogeniculate system, a classic model system for studying activity-dependent synaptic remodeling [5759], microglia were shown to eliminate synaptic connections by engulfing a subset of immature, less active presynaptic inputs [57]. Furthermore, blocking engulfment either pharmacologically or genetically through disruption of complement-mediated phagocytosis resulted in a sustained increase in synapse density and inappropriate connectivity [57, 60, 61]. These data suggest a model by which complement proteins such as C1q and C3 bind or ‘tag’ less active synapses for removal by microglia via the phagocytic receptor, complement receptor 3 (CR3). This model is supported by in vivo data that C1q and C3 localize to synaptic compartments, synaptic engulfment is reduced in C1q, C3, and CR3-deficient mice, and in vitro data that microglia clear C1q-bound neurites by CR3-dependent phagocytic signaling [6062]. It remains unknown if and how activity regulates complement proteins. Interestingly, in the context of hypoxic injury and inflammation in the hippocampus, CR3 was necessary to induce long-term synaptic depression (LTD) and suggests microglia can modulate the plasticity of functional synapses via CR3 [63]. Together, these studies demonstrate that microglia respond to changes in neural activity and suggest that they are critical to the remodeling and maturation of synaptic connections in the developing brain.

In addition to modulating development of existing connectivity, microglia have also been implicated in the initial wiring of the embryonic brain. Early work in the developing kitten corpus callosum demonstrated engulfed axonal debris within microglia and astrocytes concomitant with large-scale axonal remodeling [64]. Recent work in the embryonic mouse has demonstrated a similar phenomenon in which microglia appear to engulf a subset of developing tyrosine hydroxylase (TH)-positive, dopaminergic axons [65]. Furthermore, dopaminergic axons were increased at the entrance to the embryonic subpallium in CX3CR1-deficient mice or when mice lacked microglia due to genetic deletion of PU.1 or treatment with an antibody against colony stimulating factor 1 receptor (CSFR1) [65]. Conversely, increasing microglia activation with LPS resulted in a decrease in dopaminergic axons. Interestingly, in addition to dopaminergic axons, interneurons were also affected. Depletion or activation of microglia, as well as genetic deletion of CX3CR1 or DAP12, resulted in premature entry and abnormal distribution of Lhx6-expressing interneurons in the embryonic cortical plate and a 10% decrease in a subset of interneurons in the postnatal cortex. In another study, outgrowth and fasciculation of axons within corpus callosum were assessed in embryonic PU.1−/−, DAP12−/−, or LPS-treated mice [66]. Gene expression profiles at E17.5 revealed a down-regulation of genes related to neuritogenesis in DAP12−/− and LPS-treated mice, which were accompanied by a significant increase in defasciculated axon tracts in the corpus callosum of PU.1−/−, DAP12−/−, and LPS-treated mice. These studies suggest that impairing microglia during embryogenesis affects axon outgrowth and fasciculation.

In summary, these studies suggest key roles for microglia in the formation and remodeling of neural circuits throughout several regions of the brain (Figure 2, Key Figure). Future work is necessary to elucidate more mechanisms underlying these intercellular interactions and to identify functional consequences. For example, while microglia engulf synapses through the classical complement cascade in the developing visual system, this is likely not the only mechanism. In fact, mammalian astrocytes and Drosophila glial cells perform similar functions through different phagocytic receptors including MEGF10 and MERTK in mammals and Draper (the MEGF10 homologue) in Drosophila [6770]. These data raise the question of whether microglia and astrocytes work cooperatively. In addition, it is unknown how mechanisms that regulate microglial responses to changes in neural activity, such as NMDAR-mediated ATP signaling, regulate plasticity and maturation of circuits [47]. It is also unknown whether microglia have a preference for affecting outgrowth, synapse remodeling, and synapse maturation at specific circuits or whether this is a more global process that occurs throughout the brain. Addressing these questions will be important for understanding the basic biology underlying neural circuit development with tremendous promise for elucidating etiologies of devastating neuropsychiatric disorders with known defects in microglia and brain wiring [2, 71].

Microglia-dependent development of functional brain circuits

Data reveal new roles for microglia in sculpting structural CNS circuitry during development by regulating numbers of cells and synaptic connections as well as regulating spatial patterning of neurons and their projections. Do these functions ultimately translate to the development of functional circuits and appropriate behaviors?

Some of the most compelling evidence that microglia regulate overall circuit function and behavior are experiments in mice in which microglia have been manipulated pharmacologically or genetically. For example, infecting postnatal rats with Escherichia coli followed by a later life immune challenge with LPS resulted in increased hippocampal microglia reactivity and impaired memory in adult mice. [72] Furthermore, 15 min daily handling of postnatal pups (P4-P20) or pharmacological blockade of IL-1β, which is highly expressed by microglia in this context, attenuated these effects [73, 74]. Similarly, immune challenge in a pregnant mouse or nonhuman primate results in offspring with behavioral deficits associated with autism such as changes in ultrasonic vocalizations, abnormal social interactions, and increased repetitive behaviors [7577]. A similar prenatal immune challenge followed by peripubertal stress in the offspring also resulted in increased microglial reactivity in the pubescent hippocampus and behavioral abnormalities in adult offspring including sensorimotor gating deficits and hypersensitivity to psychotomimetic drugs [78]. In addition, another study manipulated microglia function with minocycline and observed changes in baseline, sex-specific behaviors and synapse architecture [79].

While these studies suggest that microglia can regulate synaptic function, which can ultimately translate to behavior, the pharmacological agents used in these studies are not specific and affect other cells inside and outside the CNS. As a result, other work has taken advantage of powerful molecular genetic approaches to assess the role of microglia in nervous system function. For example, genetic deletion of CX3CR1, a receptor highly enriched in microglia but also expressed by other myeloid-derived cells [2224], revealed abnormalities in structural connectivity and social behaviors in adult mice [80]. In addition, re-expression of wild-type homeobox B8 (Hoxb8) in myeloid derived cells, attenuated pathological grooming behavior in Hoxb8 mutant mice [81]. Similarly, re-expression of methyl CpG binding protein 2 (Mecp2) in microglia as well as other myeloid-derived cells attenuated phenotypes in a mouse model of Rett Syndrome (Mecp2 null mice), an X-linked neurodevelopmental disorder [82, 83] (Table 1). However, these data remain controversial [84]. To more specifically manipulate microglia function, two groups recently created mice expressing Cre-ERT2 (Cre recombinase fused to the estrogen receptor for temporal control) under the control of CX3CR1 [85, 86]. This system takes advantage of the relatively high and stable expression of CX3CR1 in microglia and low rate of microglia turnover as compared to other CX3CR1-positive peripheral immune cells [85, 87]. Using this technology, it has been demonstrated that depleting microglia in the juvenile and early adult CNS using a diphtheria toxin strategy or ablating microglia-derived BDNF results in abnormalities in motor learning [85]. In contrast, depleting microglia using a newly, developed pharmacological strategy, has yielded conflicting results [4]. CSF1R is a cell surface receptor regulating survival, proliferation, and differentiation of microglia and other mononuclear phagocytes [2, 3]. Administration of a drug that inhibits CSF1R (PLX3397) to adult mice results in depletion of primarily microglia with little effect on behavioral measures of anxiety, motor function, learning, or memory [4]. One intriguing notion exists that depletion strongly relies on context. Perhaps, microglia are most critical for establishing brain connectivity and cytoarchitecture necessary for appropriate behaviors in development, a function less critical in the adult. Future work is necessary to identify the relative importance of these cells for overall nervous system function throughout the lifespan of the animal.

While rodents are powerful experimental models that can be used to dissect cellular and molecular mechanisms and assess intermediate phenotypes associated with human neurological disesase, [8890], there are limits to the system. We are lacking mouse models that closely mimic neurological disease, particularly those that recapitulate the range of behavioral abnormalities associated with psychiatric disorders. Thus, analysis of microglia function in humans is a necessity. Indeed, some of the first evidence suggesting that microglia may play fundamental roles in the functional development of circuits was observed in psychiatric disorders, many of which are now thought to have developmental underpinnings [71, 91, 92]. Early work in postmortem human tissue demonstrated abnormally reactive microglia in brain regions relevant to behaviors associated with a range of psychiatric disorders such as autism, schizophrenia, and bipolar disorder. For example, a study in cerebral cortex demonstrated an increase in MHC class II, human leukocyte antigen-DR (HLA-DR) immunoreactive microglia in autistic versus age-matched control patients [93]. These data suggest that there is increased microglial reactivity in the autistic brain. Since these early studies, we can now map genes to a particular disease. This capability has provided new insight into roles for microglia in nervous system function and has led to the identification of mutations in microglial genes underlying neurological disease [94] (Table 1). Included in these diseases is hereditary diffuse leukoencephalopathy with spheroids (HDLS), an autosomal dominant disease of the CNS white matter caused by mutations in the microglial surface receptor CSF1R [95]. Patients suffering from these mutations have demyelination and axonal spheroids accompanied by mood, social, cognitive, and motor impairments. In addition, loss of function mutations in the microglial surface receptors DAP12 and TREM2 cause polycystic lipomembranous osteodysplasia with sclerosing leukoencephalopathy (PLOSL; Nasu-Hakola disease) [96]. This disease is characterized by the development of psychosis and early onset, progressive dementia as well as bone cysts, which are likely due to loss of receptor function in other myeloid-derived cells. Interestingly, while symptoms typically manifest in adulthood in all these disorders, CSF1R, DAP12, and TREM2 are expressed in microglia throughout development. Thus, it is possible that impairments have a developmental underpinning, which become progressively worse and manifest in behavioral changes later in life. In addition to these ‘microgliopathies’, microglia-related genes have recently been identified as risk factors for a number of other neurological diseases including CD33 and TREM2 in Alzheimer’s disease, TREM2 in frontotemporal dementia, TNFRSF1A and IRF8 in multiple sclerosis, and myeloid cell receptor P2RX7 in bipolar and major depressive disorders [97104]. In a very recent and exciting study, allelic variations in complement component 4 (C4) in humans were identified as risk factors for developing schizophrenia [105]. Furthermore, human C4 localized to synaptic compartments and mice deficient in C4 had sustained deficits in synaptic remodeling. The authors proposed that similar to C1q and C3, C4 may regulate micorglia-dependent synaptic remodeling. Future work is necessary to determine if these mutations or allelic variations are causative [106].

Concluding Remarks

It is an exciting time to study microglia (see Outstanding Questions). There are now interesting data that microglia can perform a variety of functions in the context of the developing brain including regulating number and maturation of other resident CNS cell types, vascular branching, sculpting synaptic connectivity, regulating axon outgrowth, modulating synaptic maturation and affecting overall behavior (Figure 2, Key Figure). Despite the flurry of new data, many in vitro experiments still require in vivo validation and many studies have used non-specific pharmacological approaches to study microglia function. While recent work has made exciting progress identifying roles for microglia-specific molecules in brain wiring and function, molecular mechanisms are still lacking. Furthermore, data suggest that microglia have separable functions in different brain regions, but elucidating how these regional differences are specified on a cellular and molecular level (i.e., microglial heterogeneity) will be important. Addressing these gaps in knowledge will require new tool development. In particular, there is a need for the identification of more microglia-specific genes that can be used to modulate function as well as the development of strategies to more acutely modulate microglial gene expression in a region-specific manner (e.g. viral-mediated gene delivery). These advancements will have tremendous impact on understanding microglia function in the healthy CNS. Finally, there are a large number of neurological disorders where microglia have now been implicated as central players in disease onset and/or progression [2, 107] (Table 1). However, it is impossible to model the full range of behavioral abnormalities characteristic of human disease in rodents, particularly in the case of psychiatric disorders. Thus, a more sophisticated assessment of microglia function in human patients through functional imaging and gene profiling offers great promise. Identifying molecular mechanisms in the context of animal models and developing technology to assess dysfunction in humans will be critical next steps. These advancements will be necessary to elucidate the basic biology underlying microglia function and for developing diagnostics and therapeutics for devastating neurological disorders with underlying microglia dysfunction.

Outstanding Questions Box.

  • Are microglia heterogenous? Microglia regulate an array of functions simultaneously with a high degree of regional specificity. Determining whether microglia are a heterogeneous cell population and identifying how this heterogeneity arises are critical future directions.

  • Do microglia regulate the development and maturation of non-neuronal CNS cell types in vivo? The majority of work identifying roles for microglia in regulating the development of non-neuronal cells is in vitro. Thus, it is necessary to determine if the same mechanisms and functions occur in vivo and regulate development throughout multiple regions of the CNS.

  • Do microglia work cooperatively with neurons and/or astrocytes to actively initiate synapse remodeling? All data demonstrating that microglia engulf synaptic elements in the developing brain has been in fixed tissue. Therefore, it is unknown whether microglia actively initiate synaptic remodeling and engulf intact synapses or whether they are more passively cleaning up synaptic remnants rendered vulnerable by other neuron or astrocyte-specific mechanisms.

  • Do mechanisms regulating neuronal development act in the same pathway or in parallel and are they activity-dependent? Several molecular mechanisms have been identified to regulate microglia-dependent development of neurons and their synaptic connections. However, it is unknown whether these molecular mechanisms work in the same pathway or in parallel or whether these mechanisms are regulated by neural activity.

  • Are microglia causative in neurological disorders? Pharmacological and genetic manipulation of microglia has demonstrated changes in behavior in mice and non-human primates. In addition, microglia-related genes have been identified as risk factors for neurological disorders ranging from Alzheimer’s disease to schizophrenia. Whether microglia in the developing brain have roles in disease etiology and behavioral abnormalities remains a mystery.

  • Microglia simultaneously regulate programmed cell death and survival of developing neurons.

  • Microglia promote differentiation and provide trophic support for developing astrocytes, oligodendrocytes, and vasculature.

  • Formation of neural circuits and activity-dependent remodeling and maturation of synapses are regulated by microglia in the embryonic and postnatal brain.

  • Pharmacological and genetic disruption of microglia function results in behavioral abnormalities in juvenile and adult animals.

  • Mutations in microglial genes have been identified in human neurological disease including hereditary diffuse leukoencephalopathy with spheroids (HDLS) and Nasu-Hakola disease. In addition, allelic variations in microglia-related genes are associated with increased susceptibility in a number of neurological diseases ranging from Alzheimer’s disease to schizophrenia.

Acknowledgments

Work in this publication was supported by the National Institute of Mental Health of the National Institutes of Health (R00MH102351;DPS). We thank G. Gunner, E. Mondo, and K. Johnson for their careful review and edits of this manuscript.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Ginhoux F, et al. Fate mapping analysis reveals that adult microglia derive from primitive macrophages. Science. 2010;330:841–845. doi: 10.1126/science.1194637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Prinz M, Priller J. Microglia and brain macrophages in the molecular age: from origin to neuropsychiatric disease. Nature reviews. Neuroscience. 2014;15:300–312. doi: 10.1038/nrn3722. [DOI] [PubMed] [Google Scholar]
  • 3.Ginhoux F, et al. Origin and differentiation of microglia. Frontiers in cellular neuroscience. 2013;7:45. doi: 10.3389/fncel.2013.00045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Elmore MR, et al. Colony-stimulating factor 1 receptor signaling is necessary for microglia viability, unmasking a microglia progenitor cell in the adult brain. Neuron. 2014;82:380–397. doi: 10.1016/j.neuron.2014.02.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Bruttger J, et al. Genetic Cell Ablation Reveals Clusters of Local Self-Renewing Microglia in the Mammalian Central Nervous System. Immunity. 2015;43:92–106. doi: 10.1016/j.immuni.2015.06.012. [DOI] [PubMed] [Google Scholar]
  • 6.Kriegstein A, Alvarez-Buylla A. The glial nature of embryonic and adult neural stem cells. Annual review of neuroscience. 2009;32:149–184. doi: 10.1146/annurev.neuro.051508.135600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Vaux DL, Korsmeyer SJ. Cell death in development. Cell. 1999;96:245–254. doi: 10.1016/s0092-8674(00)80564-4. [DOI] [PubMed] [Google Scholar]
  • 8.Oppenheim RW. Cell death during development of the nervous system. Annu Rev Neurosci. 1991;14:453–501. doi: 10.1146/annurev.ne.14.030191.002321. [DOI] [PubMed] [Google Scholar]
  • 9.Ferrer I, et al. Naturally occurring cell death in the cerebral cortex of the rat and removal of dead cells by transitory phagocytes. Neuroscience. 1990;39:451–458. doi: 10.1016/0306-4522(90)90281-8. [DOI] [PubMed] [Google Scholar]
  • 10.Bessis A, et al. Microglial control of neuronal death and synaptic properties. Glia. 2007;55:233–238. doi: 10.1002/glia.20459. [DOI] [PubMed] [Google Scholar]
  • 11.Frade JM, Barde YA. Microglia-derived nerve growth factor causes cell death in the developing retina. Neuron. 1998;20:35–41. doi: 10.1016/s0896-6273(00)80432-8. [DOI] [PubMed] [Google Scholar]
  • 12.Marin-Teva JL, et al. Microglia promote the death of developing Purkinje cells. Neuron. 2004;41:535–547. doi: 10.1016/s0896-6273(04)00069-8. [DOI] [PubMed] [Google Scholar]
  • 13.Sedel F, et al. Macrophage-derived tumor necrosis factor alpha, an early developmental signal for motoneuron death. J Neurosci. 2004;24:2236–2246. doi: 10.1523/JNEUROSCI.4464-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Mazaheri F, et al. Distinct roles for BAI1 and TIM-4 in the engulfment of dying neurons by microglia. Nature communications. 2014;5:4046. doi: 10.1038/ncomms5046. [DOI] [PubMed] [Google Scholar]
  • 15.Cunningham CL, et al. Microglia regulate the number of neural precursor cells in the developing cerebral cortex. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2013;33:4216–4233. doi: 10.1523/JNEUROSCI.3441-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Chamak B, et al. Brain macrophages stimulate neurite growth and regeneration by secreting thrombospondin. J Neurosci Res. 1994;38:221–233. doi: 10.1002/jnr.490380213. [DOI] [PubMed] [Google Scholar]
  • 17.Morgan SC, et al. Microglia release activators of neuronal proliferation mediated by activation of mitogen-activated protein kinase, phosphatidylinositol-3-kinase/Akt and delta-Notch signalling cascades. J Neurochem. 2004;90:89–101. doi: 10.1111/j.1471-4159.2004.02461.x. [DOI] [PubMed] [Google Scholar]
  • 18.Nagata K, et al. Microglia-derived plasminogen enhances neurite outgrowth from explant cultures of rat brain. Int J Dev Neurosci. 1993;11:227–237. doi: 10.1016/0736-5748(93)90081-n. [DOI] [PubMed] [Google Scholar]
  • 19.Antony JM, et al. Endogenous microglia regulate development of embryonic cortical precursor cells. J Neurosci Res. 2011;89:286–298. doi: 10.1002/jnr.22533. [DOI] [PubMed] [Google Scholar]
  • 20.Gebara E, et al. Adult hippocampal neurogenesis inversely correlates with microglia in conditions of voluntary running and aging. Front Neurosci. 2013;7:145. doi: 10.3389/fnins.2013.00145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Ueno M, et al. Layer V cortical neurons require microglial support for survival during postnatal development. Nat Neurosci. 2013;16:543–551. doi: 10.1038/nn.3358. [DOI] [PubMed] [Google Scholar]
  • 22.Mizutani M, et al. The fractalkine receptor but not CCR2 is present on microglia from embryonic development throughout adulthood. J Immunol. 2012;188:29–36. doi: 10.4049/jimmunol.1100421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Wolf Y, et al. Microglia, seen from the CX3CR1 angle. Frontiers in cellular neuroscience. 2013;7:26. doi: 10.3389/fncel.2013.00026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Nishiyori A, et al. Localization of fractalkine and CX3CR1 mRNAs in rat brain: does fractalkine play a role in signaling from neuron to microglia? FEBS letters. 1998;429:167–172. doi: 10.1016/s0014-5793(98)00583-3. [DOI] [PubMed] [Google Scholar]
  • 25.Shigemoto-Mogami Y, et al. Microglia enhance neurogenesis and oligodendrogenesis in the early postnatal subventricular zone. J Neurosci. 2014;34:2231–2243. doi: 10.1523/JNEUROSCI.1619-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Nakanishi M, et al. Microglia-derived interleukin-6 and leukaemia inhibitory factor promote astrocytic differentiation of neural stem/progenitor cells. The European journal of neuroscience. 2007;25:649–658. doi: 10.1111/j.1460-9568.2007.05309.x. [DOI] [PubMed] [Google Scholar]
  • 27.Lu H, et al. Optimal dose of zinc supplementation for preventing aluminum-induced neurotoxicity in rats. Neural regeneration research. 2013;8:2754–2762. doi: 10.3969/j.issn.1673-5374.2013.29.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.O'Kusky J, Ye P. Neurodevelopmental effects of insulin-like growth factor signaling. Frontiers in neuroendocrinology. 2012;33:230–251. doi: 10.1016/j.yfrne.2012.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Nicholas RS, et al. Nonactivated microglia promote oligodendrocyte precursor survival and maturation through the transcription factor NF-kappa B. The European journal of neuroscience. 2001;13:959–967. doi: 10.1046/j.0953-816x.2001.01470.x. [DOI] [PubMed] [Google Scholar]
  • 30.Clemente D, et al. The effect of glia-glia interactions on oligodendrocyte precursor cell biology during development and in demyelinating diseases. Frontiers in cellular neuroscience. 2013;7:268. doi: 10.3389/fncel.2013.00268. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Zhang X, et al. Cellular iron status influences the functional relationship between microglia and oligodendrocytes. Glia. 2006;54:795–804. doi: 10.1002/glia.20416. [DOI] [PubMed] [Google Scholar]
  • 32.Cheepsunthorn P, et al. Cellular distribution of ferritin subunits in postnatal rat brain. The Journal of comparative neurology. 1998;400:73–86. [PubMed] [Google Scholar]
  • 33.Perry VH, et al. Immunohistochemical localization of macrophages and microglia in the adult and developing mouse brain. Neuroscience. 1985;15:313–326. doi: 10.1016/0306-4522(85)90215-5. [DOI] [PubMed] [Google Scholar]
  • 34.Santos AM, et al. Embryonic and postnatal development of microglial cells in the mouse retina. The Journal of comparative neurology. 2008;506:224–239. doi: 10.1002/cne.21538. [DOI] [PubMed] [Google Scholar]
  • 35.Kubota Y, et al. M-CSF inhibition selectively targets pathological angiogenesis and lymphangiogenesis. The Journal of experimental medicine. 2009;206:1089–1102. doi: 10.1084/jem.20081605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Checchin D, et al. Potential role of microglia in retinal blood vessel formation. Investigative ophthalmology & visual science. 2006;47:3595–3602. doi: 10.1167/iovs.05-1522. [DOI] [PubMed] [Google Scholar]
  • 37.Fantin A, et al. Tissue macrophages act as cellular chaperones for vascular anastomosis downstream of VEGF-mediated endothelial tip cell induction. Blood. 2010;116:829–840. doi: 10.1182/blood-2009-12-257832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Rymo SF, et al. A two-way communication between microglial cells and angiogenic sprouts regulates angiogenesis in aortic ring cultures. PloS one. 2011;6:e15846. doi: 10.1371/journal.pone.0015846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Stefater JA, 3rd, et al. Regulation of angiogenesis by a non-canonical Wnt-Flt1 pathway in myeloid cells. Nature. 2011;474:511–515. doi: 10.1038/nature10085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Unoki N, et al. SDF-1/CXCR4 contributes to the activation of tip cells and microglia in retinal angiogenesis. Investigative ophthalmology & visual science. 2010;51:3362–3371. doi: 10.1167/iovs.09-4978. [DOI] [PubMed] [Google Scholar]
  • 41.Hua JY, Smith SJ. Neural activity and the dynamics of central nervous system development. Nature neuroscience. 2004;7:327–332. doi: 10.1038/nn1218. [DOI] [PubMed] [Google Scholar]
  • 42.Katz L, Shatz C. Synaptic activity and the constuction of cortical circuits. Science. 1996;274:1133–1138. doi: 10.1126/science.274.5290.1133. [DOI] [PubMed] [Google Scholar]
  • 43.Pocock JM, Kettenmann H. Neurotransmitter receptors on microglia. Trends in neurosciences. 2007;30:527–535. doi: 10.1016/j.tins.2007.07.007. [DOI] [PubMed] [Google Scholar]
  • 44.Schafer DP, Stevens B. Microglia Function in Central Nervous System Development and Plasticity. Cold Spring Harbor perspectives in biology. 2015;7 doi: 10.1101/cshperspect.a020545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Tremblay ME. The role of microglia at synapses in the healthy CNS: novel insights from recent imaging studies. Neuron glia biology. 2011;7:67–76. doi: 10.1017/S1740925X12000038. [DOI] [PubMed] [Google Scholar]
  • 46.Wu Y, et al. Microglia: Dynamic Mediators of Synapse Development and Plasticity. Trends in immunology. 2015;36:605–613. doi: 10.1016/j.it.2015.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Dissing-Olesen L, et al. Activation of neuronal NMDA receptors triggers transient ATP-mediated microglial process outgrowth. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2014;34:10511–10527. doi: 10.1523/JNEUROSCI.0405-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Nimmerjahn A, et al. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science. 2005;308:1314–1318. doi: 10.1126/science.1110647. [DOI] [PubMed] [Google Scholar]
  • 49.Davalos D, et al. ATP mediates rapid microglial response to local brain injury in vivo. Nature neuroscience. 2005;8:752–758. doi: 10.1038/nn1472. [DOI] [PubMed] [Google Scholar]
  • 50.Li Y, et al. Reciprocal regulation between resting microglial dynamics and neuronal activity in vivo. Developmental cell. 2012;23:1189–1202. doi: 10.1016/j.devcel.2012.10.027. [DOI] [PubMed] [Google Scholar]
  • 51.Tremblay ME, et al. Microglial interactions with synapses are modulated by visual experience. PLoS Biol. 2010;8:e1000527. doi: 10.1371/journal.pbio.1000527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Wake H, et al. Resting microglia directly monitor the functional state of synapses in vivo and determine the fate of ischemic terminals. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2009;29:3974–3980. doi: 10.1523/JNEUROSCI.4363-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Roumier A, et al. Impaired synaptic function in the microglial KARAP/DAP12-deficient mouse. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2004;24:11421–11428. doi: 10.1523/JNEUROSCI.2251-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Roumier A, et al. Prenatal activation of microglia induces delayed impairment of glutamatergic synaptic function. PLoS One. 2008;3:e2595. doi: 10.1371/journal.pone.0002595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Paolicelli RC, et al. Synaptic pruning by microglia is necessary for normal brain development. Science. 2011;333:1456–1458. doi: 10.1126/science.1202529. [DOI] [PubMed] [Google Scholar]
  • 56.Hoshiko M, et al. Deficiency of the microglial receptor CX3CR1 impairs postnatal functional development of thalamocortical synapses in the barrel cortex. J Neurosci. 2012;32:15106–15111. doi: 10.1523/JNEUROSCI.1167-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Schafer DP, et al. Microglia sculpt postnatal neural circuits in an activity and complement-dependent manner. Neuron. 2012;74:691–705. doi: 10.1016/j.neuron.2012.03.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Chen C, Regehr WG. Developmental remodeling of the retinogeniculate synapse. Neuron. 2000;28:955–966. doi: 10.1016/s0896-6273(00)00166-5. [DOI] [PubMed] [Google Scholar]
  • 59.Hong YK, Chen C. Wiring and rewiring of the retinogeniculate synapse. Current opinion in neurobiology. 2011;21:228–237. doi: 10.1016/j.conb.2011.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Stevens B, et al. The classical complement cascade mediates CNS synapse elimination. Cell. 2007;131:1164–1178. doi: 10.1016/j.cell.2007.10.036. [DOI] [PubMed] [Google Scholar]
  • 61.Bialas AR, Stevens B. TGF-beta signaling regulates neuronal C1q expression and developmental synaptic refinement. Nature neuroscience. 2013;16:1773–1782. doi: 10.1038/nn.3560. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 62.Linnartz B, et al. Sialic acid on the neuronal glycocalyx prevents complement C1 binding and complement receptor-3-mediated removal by microglia. The Journal of neuroscience : the official journal of the Society for Neuroscience. 2012;32:946–952. doi: 10.1523/JNEUROSCI.3830-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Zhang J, et al. Microglial CR3 activation triggers long-term synaptic depression in the hippocampus via NADPH oxidase. Neuron. 2014;82:195–207. doi: 10.1016/j.neuron.2014.01.043. [DOI] [PubMed] [Google Scholar]
  • 64.Berbel P, Innocenti GM. The development of the corpus callosum in cats: a light- and electron-microscopic study. The Journal of comparative neurology. 1988;276:132–156. doi: 10.1002/cne.902760109. [DOI] [PubMed] [Google Scholar]
  • 65.Squarzoni P, et al. Microglia modulate wiring of the embryonic forebrain. Cell Rep. 2014;8:1271–1279. doi: 10.1016/j.celrep.2014.07.042. [DOI] [PubMed] [Google Scholar]
  • 66.Pont-Lezica L, et al. Microglia shape corpus callosum axon tract fasciculation: functional impact of prenatal inflammation. Eur J Neurosci. 2014;39:1551–1557. doi: 10.1111/ejn.12508. [DOI] [PubMed] [Google Scholar]
  • 67.Chung WS, et al. Astrocytes mediate synapse elimination through MEGF10 and MERTK pathways. Nature. 2013;504:394–400. doi: 10.1038/nature12776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.MacDonald JM, et al. The Drosophila cell corpse engulfment receptor Draper mediates glial clearance of severed axons. Neuron. 2006;50:869–881. doi: 10.1016/j.neuron.2006.04.028. [DOI] [PubMed] [Google Scholar]
  • 69.Hoopfer ED, et al. Wlds protection distinguishes axon degeneration following injury from naturally occurring developmental pruning. Neuron. 2006;50:883–895. doi: 10.1016/j.neuron.2006.05.013. [DOI] [PubMed] [Google Scholar]
  • 70.Awasaki T, et al. Essential role of the apoptotic cell engulfment genes draper and ced-6 in programmed axon pruning during Drosophila metamorphosis. Neuron. 2006;50:855–867. doi: 10.1016/j.neuron.2006.04.027. [DOI] [PubMed] [Google Scholar]
  • 71.Frick LR, et al. Microglial dysregulation in psychiatric disease. Clinical & developmental immunology. 2013;2013:608654. doi: 10.1155/2013/608654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Bilbo SD, et al. Neonatal infection induces memory impairments following an immune challenge in adulthood. Behav Neurosci. 2005;119:293–301. doi: 10.1037/0735-7044.119.1.293. [DOI] [PubMed] [Google Scholar]
  • 73.Bilbo SD, et al. Differential effects of neonatal handling on early life infection-induced alterations in cognition in adulthood. Brain Behav Immun. 2007;21:332–342. doi: 10.1016/j.bbi.2006.10.005. [DOI] [PubMed] [Google Scholar]
  • 74.Williamson LL, et al. Microglia and memory: modulation by early-life infection. J Neurosci. 2011;31:15511–15521. doi: 10.1523/JNEUROSCI.3688-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Bauman MD, et al. Activation of the maternal immune system during pregnancy alters behavioral development of rhesus monkey offspring. Biological psychiatry. 2014;75:332–341. doi: 10.1016/j.biopsych.2013.06.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Machado CJ, et al. Maternal immune activation in nonhuman primates alters social attention in juvenile offspring. Biological psychiatry. 2015;77:823–832. doi: 10.1016/j.biopsych.2014.07.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Malkova NV, et al. Maternal immune activation yields offspring displaying mouse versions of the three core symptoms of autism. Brain, behavior, and immunity. 2012;26:607–616. doi: 10.1016/j.bbi.2012.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Giovanoli S, et al. Stress in puberty unmasks latent neuropathological consequences of prenatal immune activation in mice. Science. 2013;339:1095–1099. doi: 10.1126/science.1228261. [DOI] [PubMed] [Google Scholar]
  • 79.Lenz KM, et al. Microglia are essential to masculinization of brain and behavior. J Neurosci. 2013;33:2761–2772. doi: 10.1523/JNEUROSCI.1268-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Zhan Y, et al. Deficient neuron-microglia signaling results in impaired functional brain connectivity and social behavior. Nat Neurosci. 2014;17:400–406. doi: 10.1038/nn.3641. [DOI] [PubMed] [Google Scholar]
  • 81.Chen SK, et al. Hematopoietic origin of pathological grooming in Hoxb8 mutant mice. Cell. 2010;141:775–785. doi: 10.1016/j.cell.2010.03.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Derecki NC, et al. Wild-type microglia arrest pathology in a mouse model of Rett syndrome. Nature. 2012;484:105–109. doi: 10.1038/nature10907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Cronk JC, et al. Methyl-CpG Binding Protein 2 Regulates Microglia and Macrophage Gene Expression in Response to Inflammatory Stimuli. Immunity. 2015;42:679–691. doi: 10.1016/j.immuni.2015.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Wang J, et al. Wild-type microglia do not reverse pathology in mouse models of Rett syndrome. Nature. 2015;521:E1–E4. doi: 10.1038/nature14444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Parkhurst CN, et al. Microglia promote learning-dependent synapse formation through brain-derived neurotrophic factor. Cell. 2013;155:1596–1609. doi: 10.1016/j.cell.2013.11.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Yona S, et al. Fate mapping reveals origins and dynamics of monocytes and tissue macrophages under homeostasis. Immunity. 2013;38:79–91. doi: 10.1016/j.immuni.2012.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Goldmann T, et al. A new type of microglia gene targeting shows TAK1 to be pivotal in CNS autoimmune inflammation. Nature neuroscience. 2013;16:1618–1626. doi: 10.1038/nn.3531. [DOI] [PubMed] [Google Scholar]
  • 88.Nestler EJ, Hyman SE. Animal models of neuropsychiatric disorders. Nature neuroscience. 2010;13:1161–1169. doi: 10.1038/nn.2647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Chadman KK, et al. Criteria for validating mouse models of psychiatric diseases. American journal of medical genetics. Part B, Neuropsychiatric genetics : the official publication of the International Society of Psychiatric Genetics. 2009;150B:1–11. doi: 10.1002/ajmg.b.30777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Seong E, et al. Mouse models for psychiatric disorders. Trends in genetics : TIG. 2002;18:643–650. doi: 10.1016/s0168-9525(02)02807-x. [DOI] [PubMed] [Google Scholar]
  • 91.Beumer W, et al. The immune theory of psychiatric diseases: a key role for activated microglia and circulating monocytes. Journal of leukocyte biology. 2012;92:959–975. doi: 10.1189/jlb.0212100. [DOI] [PubMed] [Google Scholar]
  • 92.Blank T, Prinz M. Microglia as modulators of cognition and neuropsychiatric disorders. Glia. 2013;61:62–70. doi: 10.1002/glia.22372. [DOI] [PubMed] [Google Scholar]
  • 93.Vargas DL, et al. Neuroglial activation and neuroinflammation in the brain of patients with autism. Ann Neurol. 2005;57:67–81. doi: 10.1002/ana.20315. [DOI] [PubMed] [Google Scholar]
  • 94.Biber K, et al. Central nervous system myeloid cells as drug targets: current status and translational challenges. Nature reviews. Drug discovery. 2016;15:110–124. doi: 10.1038/nrd.2015.14. [DOI] [PubMed] [Google Scholar]
  • 95.Rademakers R, et al. Mutations in the colony stimulating factor 1 receptor (CSF1R) gene cause hereditary diffuse leukoencephalopathy with spheroids. Nature genetics. 2012;44:200–205. doi: 10.1038/ng.1027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Paloneva J, et al. Mutations in two genes encoding different subunits of a receptor signaling complex result in an identical disease phenotype. American journal of human genetics. 2002;71:656–662. doi: 10.1086/342259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Bradshaw EM, et al. CD33 Alzheimer's disease locus: altered monocyte function and amyloid biology. Nature neuroscience. 2013;16:848–850. doi: 10.1038/nn.3435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Griciuc A, et al. Alzheimer's disease risk gene CD33 inhibits microglial uptake of amyloid beta. Neuron. 2013;78:631–643. doi: 10.1016/j.neuron.2013.04.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Guerreiro R, et al. TREM2 variants in Alzheimer's disease. The New England journal of medicine. 2013;368:117–127. doi: 10.1056/NEJMoa1211851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Guerreiro RJ, et al. Using exome sequencing to reveal mutations in TREM2 presenting as a frontotemporal dementia-like syndrome without bone involvement. JAMA neurology. 2013;70:78–84. doi: 10.1001/jamaneurol.2013.579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.The genetic association of variants in CD6, TNFRSF1A and IRF8 to multiple sclerosis: a multicenter case-control study. PloS one. 2011;6:e18813. doi: 10.1371/journal.pone.0018813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.De Jager PL, et al. Meta-analysis of genome scans and replication identify CD6, IRF8 and TNFRSF1A as new multiple sclerosis susceptibility loci. Nature genetics. 2009;41:776–782. doi: 10.1038/ng.401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Barden N, et al. Analysis of single nucleotide polymorphisms in genes in the chromosome 12Q24.31 region points to P2RX7 as a susceptibility gene to bipolar affective disorder. American journal of medical genetics. Part B, Neuropsychiatric genetics : the official publication of the International Society of Psychiatric Genetics. 2006;141B:374–382. doi: 10.1002/ajmg.b.30303. [DOI] [PubMed] [Google Scholar]
  • 104.Lucae S, et al. P2RX7, a gene coding for a purinergic ligand-gated ion channel, is associated with major depressive disorder. Human molecular genetics. 2006;15:2438–2445. doi: 10.1093/hmg/ddl166. [DOI] [PubMed] [Google Scholar]
  • 105.Sekar A, et al. Schizophrenia risk from complex variation of complement component 4. Nature. 2016 doi: 10.1038/nature16549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Chung WS, et al. Do glia drive synaptic and cognitive impairment in disease? Nature neuroscience. 2015;18:1539–1545. doi: 10.1038/nn.4142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Ransohoff RM, Perry VH. Microglial physiology: unique stimuli, specialized responses. Annu Rev Immunol. 2009;27:119–145. doi: 10.1146/annurev.immunol.021908.132528. [DOI] [PubMed] [Google Scholar]

RESOURCES