PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray...

16
Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019 SCIENCE ADVANCES | RESEARCH ARTICLE 1 of 15 PSYCHOLOGY Corticosteroid signaling at the brain-immune interface impedes coping with severe psychological stress A. Kertser 1 * , K. Baruch 1 * , A. Deczkowska 2 , A. Weiner 2 , T. Croese 3 , M. Kenigsbuch 1 , I. Cooper 4 , M. Tsoory 5 , S. Ben-Hamo 1 , I. Amit 2 , M. Schwartz 1‡ The immune system supports brain plasticity and homeostasis, yet it is prone to changes following psychological stress. Thus, it remains unclear whether and how stress-induced immune alterations contribute to the development of mental pathologies. Here, we show that following severe stress in mice, leukocyte trafficking through the choroid plexus (CP), a compartment that mediates physiological immune-brain communication, is impaired. Blocking glucocorticoid receptor signaling, either systemically or locally through its genetic knockdown at the CP, facilitated the recruitment of Gata3- and Foxp3-expressing T cells to the brain and attenuated post-traumatic behavioral deficits. These findings functionally link post-traumatic stress behavior with elevated stress-related corticosteroid signaling at the brain-immune interface and suggest a novel therapeutic target to attenuate the consequences of severe psychological stress. INTRODUCTION Psychological stress is a common life event, which often affects brain function and cognitive performance. Severely stressful conditions can have a long-term effect on the brain and contribute to a variety of chronic diseases, ranging from depression to post-traumatic stress disorder (PTSD) (1). When changes in the environment are perceived as a threat to homeostasis that may challenge the organism’s well- being, the body responds by initiating a protective mechanism through a cascade of events, which culminates with the release of glucocorticoids from the adrenal cortex (2). Such a response, if persists, can cause sup- pressive effects on peripheral immune responses (34). This, together with the fact that adaptive immunity plays a role in the maintenance and functional plasticity of the central nervous system (CNS) (57), including in experimental animal models of coping with stress (811), emphasize that mental stress is reciprocally intertwined with immune system activity, although the extent of this relationship is not fully understood. The cross talk between the immune system and the brain is tightly regulated by the restrictive nature of the CNS anatomy and is thus confined to specialized interfaces, such as the brain’s choroid plexus (CP) and the meningeal spaces (1213). Specifically, the CP was shown to function as an immunological interface through which blood-borne leukocytes can enter the CNS territory both for immuno- surveillance in homeostasis and in the cases of brain pathologies, in which it was demonstrated that circulating immune cells contribute to brain repair (1417). Recently, CP-mediated interaction of immune cells with the CNS was shown to dysfunction in aging and under certain neurodegenerative conditions, partly due to peripheral im- mune changes (1819) and/or due to intrinsic signals from within the CNS (7). Furthermore, augmenting leukocyte trafficking across the CP by systemic immunomodulation was functionally linked with a beneficial effect on brain pathology (71922). On the basis of these findings, we hypothesized that the fate of the CP as an immunolog- ical interface that supports leukocyte trafficking and thereby brain immunosurveillance might have a critical role in mediating the ef- fects of psychological stress on anxiety-related behaviors. Here, we found that severe stress in mice results in reduced immuno- surveillance within the cerebrospinal fluid (CSF), which we found to be associated with corticosteroid (CORT) signaling at the brain- immune interface. Partial elimination of glucocorticoid receptor (GR) expression by CP epithelial cells by genetic manipulation enhanced its gateway activity for leukocyte trafficking during the stress response and promoted an anxiolytic effect, which was accompanied by ac- cumulation of nonpathogenic Gata3 and Foxp3 T cells in the brain. RESULTS Severe psychological stress negatively affects CP gateway activity On the basis of past findings showing that peripheral immune cells play an important role in stress resilience (101123), we first assessed whether severe psychological stress affects spontaneous cell traf- ficking to the CNS. To this end, we adopted a mouse model of stress that leads to sustained behavioral deficits (24). In this model, mice are subjected to two sessions of inescapable electrical foot shocks, where the first session serves as a severe stressful event (“Trauma”) and the second as a boosting trigger (“Trigger”) (Fig. 1A). Mice ex- posed to such manipulation were previously shown to exhibit hyper- vigilant behavior, impaired attention, increased risk assessment, and sleep disorders (24). Quantitative polymerase chain reaction (qPCR) analysis of the CP, isolated from the brain ventricles at different time points following stress induction, revealed a sustained decrease in mRNA levels of genes encoding for trafficking molecules, such as icam1, cxcl10, and ccl2, following the combination of Trauma and Trigger, but not Trauma alone (Fig. 1B). These trafficking-related molecules were previously demonstrated to be essential for trans- epithelial migration of leukocytes across the CP (172526), and their reduced expression levels were associated with decreased immuno- surveillance of the CNS (71920). 1 Department of Neurobiology, Weizmann Institute of Science, Rehovot, Israel. 2 Department of Immunology, Weizmann Institute of Science, Rehovot, Israel. 3 Clinical Neuroimmunology Unit, Institute of Experimental Neurology, Division of Neuroscience, San Raffaele Scientific Institute, Milan, Italy. 4 The Joseph Sagol Neuroscience Center, Sheba Medical Center, Ramat Gan, Israel. 5 Department of Veterinary Resources, Weizmann Institute of Science, Rehovot, Israel. *These authors contributed equally to this work. †Present address: ImmunoBrain Checkpoint Ltd., Ness Ziona 7414002, Israel. ‡Corresponding author. Email: [email protected] Copyright © 2019 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC). on November 17, 2020 http://advances.sciencemag.org/ Downloaded from

Transcript of PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray...

Page 1: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

1 of 15

P S Y C H O L O G Y

Corticosteroid signaling at the brain-immune interface impedes coping with severe psychological stressA. Kertser1*†, K. Baruch1*†, A. Deczkowska2, A. Weiner2, T. Croese3, M. Kenigsbuch1, I. Cooper4, M. Tsoory5, S. Ben-Hamo1, I. Amit2, M. Schwartz1‡

The immune system supports brain plasticity and homeostasis, yet it is prone to changes following psychological stress. Thus, it remains unclear whether and how stress-induced immune alterations contribute to the development of mental pathologies. Here, we show that following severe stress in mice, leukocyte trafficking through the choroid plexus (CP), a compartment that mediates physiological immune-brain communication, is impaired. Blocking glucocorticoid receptor signaling, either systemically or locally through its genetic knockdown at the CP, facilitated the recruitment of Gata3- and Foxp3-expressing T cells to the brain and attenuated post-traumatic behavioral deficits. These findings functionally link post-traumatic stress behavior with elevated stress-related corticosteroid signaling at the brain-immune interface and suggest a novel therapeutic target to attenuate the consequences of severe psychological stress.

INTRODUCTIONPsychological stress is a common life event, which often affects brain function and cognitive performance. Severely stressful conditions can have a long-term effect on the brain and contribute to a variety of chronic diseases, ranging from depression to post-traumatic stress disorder (PTSD) (1). When changes in the environment are perceived as a threat to homeostasis that may challenge the organism’s well- being, the body responds by initiating a protective mechanism through a cascade of events, which culminates with the release of glucocorticoids from the adrenal cortex (2). Such a response, if persists, can cause sup-pressive effects on peripheral immune responses (3, 4). This, together with the fact that adaptive immunity plays a role in the maintenance and functional plasticity of the central nervous system (CNS) (5–7), including in experimental animal models of coping with stress (8–11), emphasize that mental stress is reciprocally intertwined with immune system activity, although the extent of this relationship is not fully understood.

The cross talk between the immune system and the brain is tightly regulated by the restrictive nature of the CNS anatomy and is thus confined to specialized interfaces, such as the brain’s choroid plexus (CP) and the meningeal spaces (12, 13). Specifically, the CP was shown to function as an immunological interface through which blood-borne leukocytes can enter the CNS territory both for immuno-surveillance in homeostasis and in the cases of brain pathologies, in which it was demonstrated that circulating immune cells contribute to brain repair (14–17). Recently, CP-mediated interaction of immune cells with the CNS was shown to dysfunction in aging and under certain neurodegenerative conditions, partly due to peripheral im-mune changes (18, 19) and/or due to intrinsic signals from within the CNS (7). Furthermore, augmenting leukocyte trafficking across

the CP by systemic immunomodulation was functionally linked with a beneficial effect on brain pathology (7, 19–22). On the basis of these findings, we hypothesized that the fate of the CP as an immunolog-ical interface that supports leukocyte trafficking and thereby brain immunosurveillance might have a critical role in mediating the ef-fects of psychological stress on anxiety-related behaviors.

Here, we found that severe stress in mice results in reduced immuno-surveillance within the cerebrospinal fluid (CSF), which we found to be associated with corticosteroid (CORT) signaling at the brain- immune interface. Partial elimination of glucocorticoid receptor (GR) expression by CP epithelial cells by genetic manipulation enhanced its gateway activity for leukocyte trafficking during the stress response and promoted an anxiolytic effect, which was accompanied by ac-cumulation of nonpathogenic Gata3 and Foxp3 T cells in the brain.

RESULTSSevere psychological stress negatively affects CP gateway activityOn the basis of past findings showing that peripheral immune cells play an important role in stress resilience (10, 11, 23), we first assessed whether severe psychological stress affects spontaneous cell traf-ficking to the CNS. To this end, we adopted a mouse model of stress that leads to sustained behavioral deficits (24). In this model, mice are subjected to two sessions of inescapable electrical foot shocks, where the first session serves as a severe stressful event (“Trauma”) and the second as a boosting trigger (“Trigger”) (Fig. 1A). Mice ex-posed to such manipulation were previously shown to exhibit hyper-vigilant behavior, impaired attention, increased risk assessment, and sleep disorders (24). Quantitative polymerase chain reaction (qPCR) analysis of the CP, isolated from the brain ventricles at different time points following stress induction, revealed a sustained decrease in mRNA levels of genes encoding for trafficking molecules, such as icam1, cxcl10, and ccl2, following the combination of Trauma and Trigger, but not Trauma alone (Fig. 1B). These trafficking-related molecules were previously demonstrated to be essential for trans-epithelial migration of leukocytes across the CP (17, 25, 26), and their reduced expression levels were associated with decreased immuno-surveillance of the CNS (7, 19, 20).

1Department of Neurobiology, Weizmann Institute of Science, Rehovot, Israel. 2Department of Immunology, Weizmann Institute of Science, Rehovot, Israel. 3Clinical Neuroimmunology Unit, Institute of Experimental Neurology, Division of Neuroscience, San Raffaele Scientific Institute, Milan, Italy. 4The Joseph Sagol Neuroscience Center, Sheba Medical Center, Ramat Gan, Israel. 5Department of Veterinary Resources, Weizmann Institute of Science, Rehovot, Israel.*These authors contributed equally to this work.†Present address: ImmunoBrain Checkpoint Ltd., Ness Ziona 7414002, Israel.‡Corresponding author. Email: [email protected]

Copyright © 2019 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC).

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 2: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

2 of 15

To further evaluate the effect of psychological stress on CNS im-munosurveillance, we analyzed the cellular composition of the CSF using flow cytometry. We found a transient reduction in CD45+ leukocyte numbers, including both CD11b+ and CD4+ cells, occurring 24 to 48 hours following the last session of foot shocks (counting after the Trigger session in the sequential combination of the two sessions) (Fig. 1C). Analyzing the brain tissue, excluding the CP and most of the peripheral meninges, revealed a similar decrease in the numbers of CD4+ T cells (fig. S1). Further analysis revealed a positive correlation between the number of leukocytes in the CSF and gene expression levels of icam1, cxcl10, and ccl2 at the CP (Fig. 1D), substantiating the functional linkage between CP expression of genes encoding for leukocyte trafficking and CNS immunosurveillance.

Boosting the peripheral immune response before psychological stress enhances brain immunosurveillance and attenuates stress-related behavioral deficitsThe observed reduction of immunosurveillance following severe stress, coupled with the linkage between CP dysfunction and the patho-physiology of neurodegenerative diseases (7, 19), led us to suggest that boosting CP gateway activity to support leukocyte trafficking would improve the ability to cope with stress via increased immu-nosurveillance of the CNS. To test this working hypothesis, we first employed a method that was previously used to transiently eliminate peripheral regulatory T cells (Tregs), thereby activating the immune system and evoking an interferon- (IFN-)–mediated immune re-sponse, which is associated with augmented CP gateway activity for leukocyte trafficking (19). Specifically, we used Foxp3.LuciDTR mice that express a diphtheria toxin (DTx) receptor under the FoxP3 promotor (FoxP3-DTR) (27), which enables the transient depletion of Tregs upon administration of DTx (19). Notably, wild-

type mice lack DTx receptor; thus, administration of the toxin to wild-type mice in this paradigm was not found to lead to any alterations in CP activity or blood-brain barrier (BBB) integrity (19, 28, 29). FoxP3-DTR mice that were subjected to both Trauma and Trigger were injected daily with DTx for four consecutive days, starting from 2 days before the Trauma session, to achieve the depletion of Tregs at a time point coinciding with the end of stress induction, namely, right after the Trigger session (Fig. 2A and fig. S2). Analysis of gene expression 24 hours following the end of stress induction (24 hours after the Trigger session), a time point that showed a marked decrease in CP gateway activity in response to stress (Fig. 1), revealed an increase in the expression of leukocyte trafficking–related genes following Treg depletion, among which cxcl10 was the most pronounced (Fig. 2B).

The enhanced expression by the CP of genes encoding for leukocyte trafficking molecules, following the depletion of Tregs, prompted us to assess whether this manipulation could subsequently increase the homing of immune cells into the CNS parenchyma. Immunohisto-chemical analysis of brain tissues 2 weeks following stress induction revealed that transient depletion of Tregs resulted in the accumula-tion of CD3+ cells, which were mostly observed along the interven-tricular foramina adjacent to the CP and in the hippocampus (HC) (Fig. 2, C to E). Moreover, about a quarter (26 ± 4%) of the CD3+ T cells in the HC were found to be positive for FoxP3+ (fig. S3), which might indicate an outcome of local conversion of T cell phenotype from effector cells to regulatory cells (30). Further flow cytometry analysis of whole brains (excluding the CP and most of the meningeal tissues) at 2 weeks following stress induction confirmed the in-creased numbers of CD45+/TCR+ T cells in mice that were transiently depleted of Tregs (Fig. 2F). To exclude potential cell entry through a breached BBB, we examined the endothelial basal lamina in both

Fig. 1. CP gateway dysfunction for immunosurveillance following severe psychological stress. (A) Scheme depicting the experimental design, showing the sched-ule of two sessions of stress induction by electrical foot shock followed by the testing period. (B) Gene expression levels, as measured by qPCR, in CP of mice following Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before Trauma) (n = 6 to 14 per group; Student’s t test; data represent combined analysis of results obtained from seven independent experiments). A.U., arbitrary units. (C) Quantitative analysis of leukocyte numbers (total number of CD45+ leukocytes; CD4+ and CD11b+ cells out of total CD45+) in the CSF of mice following Trauma + Trigger. d, day; each dot represents number of cells per microliter CSF of a single mouse [n = 12 to 13 per group; one-way analysis of variance (ANOVA) (F = 7.808; P = 0.0016 for CD45+, F = 4.516; P = 0.182 for CD11b+, F = 5.544; P = 0.0082 for CD4+) followed by Newman-Keuls post hoc test; data represent combined analysis of five independent experiments]. (D) Correlation between mRNA levels and total number of CD45+ leukocytes in the CSF of a separate cohort of stressed mice 24 hours following stress induction. Each dot represents a single mouse. In all panels, error bars represent mean ± SEM. *P < 0.05; **P < 0.01; ***P < 0.001.

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 3: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

3 of 15

Fig. 2. Breaking peripheral immune tolerance prevents post-traumatic stress behavior. (A) Scheme depicting the experimental design, showing DTx injections super-imposed on the schedule of stress induction. IHC, immunohistochemistry. (B) qPCR analysis of CP gene expression in mice 24 hours after the last stress episode, with (DTx) or without Stress Treg depletion (n = 6 per group; Student’s t test). (C) Representative fluorescence microscopy images (side panels show an enlargement of marked area) and quantification of CD3+ (red) cells at the (D) dentate gyrus (DG) of the HC and the (E) third ventricle CP 14 days following stress induction (n = 8 per group; Student’s t test). (F) Flow cytometry quantitative analysis of CD45+/TCR+ cells/brain (excluding CP and meninges) 7 days following stress induction [n = 5 per group; one-way ANOVA (F = 7.145; P = 0.0103) followed by Newman-Keuls post hoc test]. (G) Hypervigilant behavior, measured by acoustic startle reaction time (left) and amplitude (right), of stressed mice 14 days following stress induction [n = 16 per group; one-way ANOVA (F = 7.778; P = 0.0015 for reaction time, F = 7.053; P = 0.0029 for amplitude) followed by Newman-Keuls post hoc test; data are representative of two independent experiments]. In all panels, error bars represent mean ± SEM. *P < 0.05; **P < 0.01; ***P < 0.001.

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 4: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

4 of 15

treated and untreated stressed mice and found no signs of perivascu-lar cuffing associated with BBB breakdown and endothelial leukocyte transmigration (fig. S4) (31).

Peripheral immunomodulation was previously shown to affect behavioral outcome in the context of mental stress (10, 23). To assess whether augmented immune-brain trafficking through the CP might affect stress-related behavior, we repeated the experiment and tested the mice for post-traumatic hypervigilance 2 weeks following stress induction. Vehicle-treated stressed mice exhibited an increased startle response to acoustic stimulus, as previously reported in this model (24). We found that the transient depletion of Tregs in the context of stress resulted in a startle response similar to that observed in the nonstressed littermates (Fig. 2G). Together, these findings suggest that increased CP gateway activity for leukocyte trafficking following a severe stressful experience might contribute to the enhanced ability to cope with psychological stress.

Systemic blockade of GR signaling restores CP gateway activity and mitigates stress-related behavioral deficitsThe observed results that CP gateway activity might be involved in stress resilience prompted us to further investigate which component(s) of the physiological stress response could affect the CP (Fig. 1). Psy-chological stress classically leads to the activation of the hypothalamic- pituitary-adrenal (HPA) axis, which culminates in increased secretion of CORTs (32). This response affects a broad range of peripheral tissues, mediated by the activation of GRs, which, among other functions, can suppress the activity of nuclear factor B (NFB) (33), a signaling pathway found to be relevant for CP activity in mediating transepithelial migration of leukocytes to the CNS terri-tory (18). We therefore first tested whether blood levels of CORT were elevated in this model of stress, and whether subjecting the mice to the second stressor (Trauma followed by Trigger) would maintain high CORT levels. We found that 1 hour following exposure of the mice to Trauma stimulus, serum CORT levels were significantly elevated, and that the level remained high following the subsequent subjection to Trigger stimulus on the next day. After 24 hours, the levels of serum CORT returned to their baseline (Fig. 3A).

We next wished to see whether the exposure to CORT outside the context of stress would be sufficient to induce an inhibitory effect on the CP. We envisioned that even a short exposure to CORT would have an effect on CP gateway activity, if it occurs concomitantly when a “danger signal” emerges from the brain parenchyma as a result of the stress. To this end, we used a paradigm that simulates danger signaling in the brain and induces the elevation of trafficking molecules by the CP. Specifically, mice received an intracerebroventricular injection of tumor necrosis factor (TNF) and an intraperitoneal injection of CORT with respective vehicle controls. While the administration of TNF alone, as previously observed (17, 18), markedly increased CP epithelium expression of genes encoding for leukocyte trafficking molecules, cotreatment with CORT inhibited TNF induction of icam1 and ccl2, but not of cxcl10 (Fig. 3B). The short exposure to CORT, administrated to control animals that were injected intracerebroven-tricularly with phosphate-buffered saline (PBS), appeared not to affect the expression levels of these genes, suggesting that a short exposure to exogenous CORT is enough to negate the effects of a danger signal coming from the brain, but might not be sufficient to decrease the basal expression levels of trafficking-related molecules.

We further tested whether the mere prolonged systemic elevation in CORT would be sufficient to induce an inhibitory effect on the

basal CP expression of leukocyte trafficking–related genes. To over-come the relatively short biological half-life of endogenous CORT following a single dosage (34) and to avoid potential stress induced by repeated injections, CORT was provided to the mice through drinking water, similar to a method previously reported (35). Following 2 weeks of continuous CORT administration, we found a decrease in icam1, cxcl10, and ccl2 expression levels at the CP (Fig. 3C), confirming that a long-term exposure to CORT can potentially lead to suppressive effects on basal CP gateway activity.

The results shown above led us to explore whether a pharma-cological inhibition of GR, in the context of a severe stressful event, could restore CP activity and alleviate the behavioral consequence of the stress. To this end, we systemically treated mice with the GR in-hibitor mifepristone (RU-486; 20 mg/kg) 30 min before each electrical shock session. qPCR analysis of the CP for immune cell trafficking–related gene expression at 24 hours after stress induction showed increased cxcl10, ccl2, and icam1 mRNA levels in mice treated with a GR blocker relative to vehicle-treated stressed mice (Fig. 3D). As-sessment of the cellular composition of the CSF by flow cytometry revealed an increase in TCR+/CXCR3+ cells (Fig. 3E), a cell popu-lation that binds to the chemoattractant CXCL10 and was previously shown to be associated with CNS immunosurveillance (36). Immuno-histochemical analysis of brain tissues 2 weeks following stress in-duction revealed an increase in the number of CD3+ cells in the HCs of GR-blocked mice relative to vehicle-treated stressed animals (Fig. 3F and fig. S5). Together, the presence of a higher number of CD3+ cells in the HC, along with increased numbers of leukocytes in the CSF and the elevated expression of leukocyte trafficking genes by the CP, was in line with the results obtained following a transient depletion of Tregs (Fig. 2), although the effect was somewhat less intense.

To determine whether the enhancement of leukocyte-mediated immunosurveillance of the CSF via GR blockade would be linked with an improved behavior following the stress, mice were tested in a Dark/Light exploratory and risk assessment behavioral maze, as previously used to document the consequences of stress in this mu-rine experimental model (24). Anxiety-like behavior manifested in the stressed vehicle-treated mice by reduced time spent visiting the bright chamber of the Dark/Light maze, as well as greater propor-tion of time spent engaging in risk assessment, when crossing over to the bright chamber. In contrast, systemic blockade of GR at the time of stress had an anxiolytic effect, demonstrated by exploratory behavior resembling that of the control group (Fig. 3G). Additional behavioral assessment using the open-field arena revealed that GR blockade led to a trend toward an increase in the cumulative time spent exploring the arena relative to vehicle-treated stressed mice (fig. S6).

Last, combined analysis of the histological data together with the behavioral indices obtained in the Dark/Light test per individual mouse revealed that the number of HC-infiltrating CD3+ cells nega-tively correlated with the proportion of time engaged in risk assess-ment and positively with time spent in the bright arena (Fig. 3H), suggesting the possible involvement of the infiltrating cells in the attenuation of the stress-related behavior.

CORTs can directly impair brain-immune communicationWe next assessed whether the observed effect of CORT on CP gate-way activity in vivo could be, at least in part, an outcome of a direct effect on the CP epithelium or whether it primarily reflected a sup-pressive effect on the peripheral immune system. To this end, we generated a primary culture of CP epithelial cells from C57BL/6J

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 5: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

5 of 15

mice, grown to form a confluent monolayer, and challenged it for 24 hours with TNF, as an NFB-dependent inducer of CP gateway activity (17, 18), with or without CORT. While the administration of TNF alone, as previously observed (17), markedly increased ex-pression of genes encoding leukocyte trafficking molecules, the cotreatment with CORT inhibited TNF induction of trafficking genes (Fig. 4A). Similar effects were also found in a human cell line of retinal pigment epithelium (fig. S7), a tissue with partial functional similarities to the CP epithelium.

Immunocytochemical examination of murine-cultured CP epi-thelial monolayer morphology revealed that the cotreatment with

CORT reversed the remodeling of tight junctions, which was in-duced by TNF (Fig. 4B). To further evaluate the relevance of CORT signaling in CP epithelial cells to the physiological barrier proper-ties of the blood-CSF barrier, primary CP epithelial cells were cul-tured on Transwell membrane inserts, which enabled the constant monitoring of transepithelial electric resistance (TEER) of the es-tablished monolayer, a marker for the integrity of intercellular tight junctions (37). Administration of TNF to the culture medium after the CP monolayer reached confluence caused an immediate reduction of TEER, while cotreatment with CORT completely reversed this effect, resulting in TEER even higher than that measured for the cultures

Fig. 3. Systemic blockade of GR signaling restores CP gateway activity and mitigates stress-related behavioral deficits. (A) Serum CORT levels (pg/ml), measured at 1 hour following Trauma, 1 hour following Trigger preceded by Trauma, and 24 hours following Trigger preceded by Trauma [n = 4 per group; one-way ANOVA (F = 18.47; P < 0.0001) followed by Newman-Keuls post hoc test]. (B) qPCR analysis of CP gene expression 4 hours following intracerebroventricular injection of TNF (100 ng per mouse) with or without intraperitoneal injection of hydrocortisone (125 g per mouse) [n = 4 per group; one-way ANOVA (F = 5.695; P = 0.0116 for Icam1, F = 11.56; P = 0.0007 for Cxcl10, F = 6.205; P = 0.0087 for Ccl2) followed by Newman-Keuls post hoc test]. n.s., not significant. (C) qPCR analysis of CP gene expression in mice following 2 weeks of oral administration of corticosterone (100 g/ml) via drinking water (n = 6 per group; Student’s t test). (D) qPCR analysis of CP gene expression 24 hours following stress induction in mice treated with mifepristone (RU-486; GR blocker) or vehicle [n = 6 to 12 per group; one-way ANOVA (F = 5.792; P = 0.0081 for Icam1, F = 2.535; P = 0.0979 for Cxcl10, F = 6.992; P = 0.0036 for Ccl2) followed by Newman-Keuls post hoc test]. (E) Quantitative analysis of CXCR3+ and CXCR3− CD45+/TCR+ cells within the CSF 5 days following stress induction with administration of mifepristone (RU-486; GR blocker) [n = 4 per group; one-way ANOVA (F = 0.038; P = 0.9624 for CXCR3−, F = 5.938; P = 0.0227 for CXCR3+) followed by Newman-Keuls post hoc test]. (F) Immunohistochemical analysis of CD3+ cells/mm3 in the dentate gyrus of mouse HC [n = 5 to 10 per group; one-way ANOVA (F = 6.556; P = 0.0058) followed by Newman-Keuls post hoc test]. (G) Percentage of time engaged in risk assessment (right) and total time spent exploring the lit compartment of the Dark/Light test (middle) measured for nonstressed (CTRL) or stressed mice treated with GR blocker or vehicle [n = 9 to 10 mice per group; one-way ANOVA (F = 4.665; P = 0.0182 for time, F = 5.740; P = 0.0084 for risk) followed by Newman-Keuls post hoc test]. Left: Representative heat maps (top) and tracks (bottom). (H) Correlation analysis between CD3+ cells in the HC and time spent in risk assessment (left) and in the lit arena (right) in the Dark/Light test. Each dot represents a single mouse. Data are representative of two independent experiments. In all panels, error bars represent mean ± SEM. *P < 0.05; **P < 0.01; ***P < 0.001.

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 6: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

6 of 15

treated with medium alone (Fig. 4C). Moreover, the administration of CORT alone was sufficient to increase TEER, suggesting that the effect on tight junction integrity may be independent of exogenous TNF-related signaling. Overall, the direct effects of CORT on CP epithelial cells in culture suggest that the observed effect of CORT in vivo, in addition to its systemic effect on the peripheral immune system, could be an outcome of direct effect on the CP gateway.

Tissue-specific knockdown of GRs in the CP boosts its gateway activity and reduces anxiety-like behaviorsTo determine the functional role of CP CORT signaling in mediating the consequences of severe stress, we established a genetic model for the specific knockdown of GR expression at the CP. Site-specific knockdown was achieved by administration of TAT-CRE recombi-nase to the brain ventricles of mice expressing the GR gene under a

Fig. 4. Corticosteroids can directly suppress CP gateway activity for leukocyte trafficking. (A) qPCR analysis of CP gene expression in primary CP epithelial cultures, treated with either TNF or hydrocortisone (CORT) at 100 ng/ml [n = 3 per group; one-way ANOVA (F = 547.2; P < 0.0001 for Icam1, F = 72.88; P < 0.0001 for Cxcl10, F = 235.6; P < 0.0001 for Ccl2)] followed by Newman-Keuls post hoc test; error bars represent mean ± SEM. ***P < 0.001. Data are representative of at least three independent experiments. (B) Representative microscopic images of ZO-1 (red) epithelial tight junction disruption in CP cell cultures. Boxes highlight the contours of tight junction morphology. (C) TEER measured over time in primary CP epithelial cultures treated with TNF (100 ng/ml) (red), hydrocortisone (100 ng/ml) (green), or both (orange). The x axis of the graph was adjusted to omit TEER fluctuations associated with medium changes (n = 3 per group; gray area represents SE for each group).

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 7: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

7 of 15

loxP cassette (GRloxP/loxP) (38); a similar manipulation was previously reported to enable CP-specific knockdown of a floxed gene (39). To confirm that the effect of an intracerebroventricular injection of TAT-CRE remains confined to the CP, we first verified this method in mice expressing yellow fluorescent protein (YFP) under a loxP cassette (40). The consequent recombination of genes, as detected by immunohistochemistry, was found to be restricted to the ventri-cles and occurred almost exclusively at the CP (Fig. 5A).

Next, to measure the effect of GR signaling ablation on CP gate-way activity, we injected GRloxP/loxP mice intracerebroventricularly with TAT-CRE recombinase or vehicle (PBS) and allowed them to recover from the procedure for a period of 2 weeks. qPCR analysis of the mouse CP revealed a twofold reduction in GR expression following TAT-CRE recombination, relative to PBS-injected mice, which was accompanied by elevation in trafficking-related gene expression (Fig. 5B). Immunohistochemical analysis of GRloxP/loxP mice injected with TAT-CRE confirmed the decrease of GR at the protein level, which was accompanied by an increase in the intercellular adhesion molecule–1 (ICAM-1) expression (Fig. 5C). While epithelial GR ex-pression was down-regulated following the administration of TAT-CRE, no loss of GR was detected in resident leukocytes that inhabit the CP, an observation that supports the notion that a local admin-istration of TAT-CRE has no effect on the immune system.

Having established that tissue-specific knockdown of GR at the CP is sufficient to increase its gateway activity for leukocyte traffick-ing, we then assessed the potential of a similar manipulation in the context of stress. To that end, following the 2-week recovery period after TAT-CRE or vehicle injection, mice were subjected to the stress paradigm, and their CPs were analyzed by qPCR. We found that even in the context of stress, the knockdown of GR allowed increased expression of trafficking-related genes 24 hours following stress induction (Fig. 5D).

We next measured the effects of CP-specific knockdown of GR on post-traumatic anxiety-like behaviors. We found that knock-down of the GR in the context of stress partially attenuated startle responses to acoustic stimuli (Fig. 5E), showed a trend toward in-creased time spent exploring the open arms of an elevated plus maze (Fig. 5F), and diminished anxiety behavior as indicated by both re-duced risk assessment and increased time spent in the bright arena of a Dark/Light maze (Fig. 5G). Together, these findings further substantiate the notion that CORT signaling in CP epithelial cells plays a role in modulating CP gateway activity for leukocyte trafficking from the blood circulation to the CNS territory.

Increased CP gateway activity under psychological stress leads to recruitment of immune-modulating T cells into the brainLast, we wished to determine whether GR reduction was associated with increased immune cell trafficking in the context of stress and the nature of these cells. Flow cytometry analysis of the brains of TAT-CRE and vehicle-injected stressed and naïve mice, excluding the CP and most of the meningeal tissues, identified an increase in the number of CD45+TCR+ T cells following GR knockdown. This increase was reflected both in absolute cell number and in the per-centage of T cells out of total CD45+ cells (Fig. 6, A and B). Knock-down of GR alone, in the absence of stress, was not sufficient to induce trafficking of T cells; however, in stressed mice, the recruitment of T cells correlated with the degree of GR knockdown (Fig. 6C). We further compared the number of brain-infiltrating T cells in TAT-CRE–injected mice, relative to their anxiety indices, and found that

their presence correlated with the degree of attenuation in the anxiety- like behavior (Fig. 6D), suggesting that these cells play a positive role in coping with stress.

The finding that augmenting the gateway activity of the CP in the context of stress leads to attenuated post-traumatic behavioral aberrations prompted us to further analyze the population of cells that are recruited into the brain. Because the overall number of in-filtrating immune cells was too low for further in-depth assessment by conventional flow cytometry, we opted for a more sensitive anal-ysis method using single-cell RNA sequencing (RNA-seq). To this end, the brains of GR knockdown and sham-treated stressed mice were processed into a single-cell suspension for flow cytometry, and the CD45+CD11b−TCR+ cells were sorted for single-cell RNA sequencing using massively parallel single-cell RNA sequencing (MARS-seq). Unsupervised clustering analysis of T cells from both GR knockdown and sham-treated mice into meta-cells (see Materials and Methods) revealed the presence of both CD4+ and CD8+ cells. Classification of the CD4+ cells according to the expression of tran-scription factors associated with different T helper cell subsets iden-tified these as mainly T helper 2 (TH2) cells (Gata3) and Tregs (Foxp3), with no detectable expression of either TH1 (Tbx21) or TH17 (Rorc) (Fig. 6, E to G). Comparison between the cell populations found in the brains of vehicle and TAT-CRE–injected stressed mice showed no significant differences in their meta-cell distribution profile, except for the overall number of recruited cells (Fig. 6H and fig. S10). These findings suggest that the level of CP gateway activity affects the number, but not the nature, of T cells recruited into the brain, which might indicate that entry of leukocytes into the brain is a physiological response that occurs spontaneously but is apparently insufficient under severe stress. Moreover, it supports the contention that improved recovery from the stress, achieved by reducing CORT signaling at the CP, enhanced a natural process of coping with stress.

DISCUSSIONThe present study demonstrates a functional linkage between the ability to cope with mental stress and CNS immunosurveillance, which is mediated, at least in part, via the brain’s CP. We show that under severe psychological stress, CP gateway activity for leukocyte trafficking is suppressed, whereas restoration of this activity, either by transiently breaking peripheral immune tolerance or by neutral-izing CORT signaling, recruits TH2 and Tregs to the CNS territory and attenuates anxiety-like behaviors. Notably, by demonstrating both in vitro and in vivo that CORT signaling could directly inhibit CP gateway activity, we identified a novel GR-dependent stress- induced pathophysiology of the CP that affects the ability to cope with mental stress.

The activity of the CP as a gateway for leukocyte trafficking was previously shown to be affected by signals derived from both sides of the blood-CSF barrier—apically from the CSF compartment, and basally from the CP stroma or blood circulation—rendering it sus-ceptible to rapid changes in the levels of cellular and soluble factors (7, 18, 41, 42). Therefore, the decrease in CP gateway activity, re-ported here in the context of severe mental stress, could result not only from direct GR-mediated signaling within the CP epithelium but also via CORT-mediated immune suppression of peripheral immune activity that would affect the overall availability of IFN-. Accordingly, alleviation of peripheral immune suppression by transient depletion of Tregs before exposure to the stress, as well as systemic

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 8: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

8 of 15

Fig. 5. Knockdown of GR expression at the CP increases its gateway activity and attenuates symptoms of PTSD-like behavior. (A) Representative fluorescence microscopy image of R26R-EYFP mice injected intracerebroventricularly with TAT-CRE to demonstrate that the expression of YFP is restricted to the CP. DAPI, 4′,6-diamidino- 2-phenylindole; Lat.v, lateral ventricle. (B) qPCR analysis of NR3C1 (Gr) and trafficking-related gene expression levels at the CP of GRloxP/loxP animals 2 weeks following intracerebroventricular TAT-CRE administration (n = 6 per group; Student’s t test). (C) Immunohistochemistry analysis and representative images of CPs from GRloxP/loxP mice after TAT-CRE injection. Images show ICAM-1 (green, top left) and GR (red, top right), merged with nuclear staining (blue, bottom left), and an enlargement of the marked area (bottom right). Arrowheads show epithelial cells expressing decreased levels of GR (n = 2 animals per group; Student’s t test). (D) CP gene expression 24 hours following stress induction in GRloxP/loxP mice pre-injected with TAT-CRE or vehicle (n = 6 per group; Student’s t test). (E) Acoustic startle reaction time (left) and amplitude (right) of stressed mice 14 days following stress induction (n = 6 to 7 animals per group; Student’s t test). (F) Time spent exploring the open arms of an elevated plus maze 12 days following stress induction (n = 6 to 7 animals per group; Student’s t test). (G) Percentage of time mice engaged in risk assessment (middle) and total time spent exploring the lit compartment of the Dark/Light maze (right) 7 days following stress induction (n = 5 per group; Student’s t test; representative graph of two independent experiments). Left: Representative heat maps (top) and tracks (bottom). In all panels, error bars represent mean ± SEM. *P < 0.05; **P < 0.01; ***P < 0.001.

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 9: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

9 of 15

Fig. 6. Increased CP gateway activity under psychological stress leads to recruitment of nonpathological T cells into the brain. (A) Flow cytometry of T cell (CD45+CD11b−TCR+) ratio out of total CD45+ cells in the brains (excluding CP and most of the meningeal tissues) of TAT-CRE or vehicle intracerebroventricularly injected stressed and unstressed (naïve) GRloxP/loxP mice 17 days following stress induction [n = 3 to 7 mice per group; one-way ANOVA (F = 6.170; P = 0.0061) followed by Newman-Keuls post hoc test]. (B) Flow cytometry quantification of CD45+/CD4+ T cells in the brains (excluding CP and meningeal tissues) of TAT-CRE or vehicle-injected mice at day 14 following stress (n = 5 per group; Student’s t test). (C) Correlation analysis between brain T cell to microglia ratio and the relative expression of the GR in TAT-CRE (black) and vehicle (red) intracerebroventricularly injected stressed GRloxP/loxP mice. (D) Correlation between the number of CD4+ cells in the brain and time spent in risk assess-ment (left) and in the lit arena (right) of the Dark/Light maze (red, vehicle treated; black, TAT-CRE intracerebroventricularly injected GRloxP/loxP mice). Statistical analysis of the curve fit was measured only for the TAT-CRE group, excluding the PBS-treated group. Each dot represents a single mouse. (E) Heat map showing the average expression of selected genes by 1600 brain-infiltrating CD45+CD11−bTCR+ cells, clustered into 18 meta-cells. Meta-cells are ordered using hierarchical clustering of T cell–specific gene lists. Colors represent a spectrum of each row relative minimal (blue) to maximal (red) gene expression. (F) Proportion of T cells classified as CD4 or CD8 according to their gene expression pattern. (G) Assessed ratios between classical helper T cell phenotypes, as derived by the product of transcription factor expression and the number of cells in each cluster. (H) Abundance of each meta-cell cluster (calculated as the average unique molecular identifiers per cell in each meta-cell) in stressed, vehicle (sham, red)–injected, or TAT-CRE (CRE, black)–injected mice. In all panels, error bars represent mean ± SEM. *P < 0.05; **P < 0.01; ***P < 0.001.

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 10: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

10 of 15

GR blockade, prevented CP dysfunction and partially attenuated the stress-induced increase in anxiety-like behaviors. We found that the mere blocking of CORT signaling at the CP, without any systemic immunomodulation, was sufficient to attenuate the stress-induced increase in anxiety-like behaviors, a finding that emphasizes the direct involvement of CP epithelium in coping with psychological stress.

In this study, the CPs from all four brain ventricles were taken together as a uniform functional cohort; however, it is possible that CP tissues from different ventricles have distinct immunological functions and different susceptibility to the effects of CORT. In addition, other roles of the CP, which are not related to leukocyte trafficking, were not investigated in this work. It is thus possible that CORT signaling might affect functions of the CP unrelated to leukocyte trafficking, such as CSF formation and clearance.

The engagement of CORTs with their receptor (GR) can form a transrepressive complex with NFB, which prevents it from binding to its target genes (43). Recently, the NFB signaling pathway in the CP epithelium was shown to be essential for the induction of leukocyte trafficking activity (18). Accordingly, the potent ability of CORTs to antagonize TNF-induced gene expression and enhance tight junction integrity in primary CP epithelial cultures supports the possibility that sustained stress-associated GR signaling could render the CP unresponsive to pathology-derived CNS danger signals. Glucocor-ticoid secretion in the organism is constantly maintained at a basal level that fluctuates over the course of the day. Nevertheless, in our experience, we could not detect any changes in CP gateway activity when comparing the expression of trafficking-related molecules at different times of the circadian cycle.

The GR, in addition to its transrepression activities, can serve as a transcription factor, with a wide array of transactivation targets that can lead to diverse effects throughout many organs of the body. One of the suspected functions of CORT that extends beyond immune- mediated functions is neuronal memory consolidation, which was suggested to be affected during the stress response (44). Accordingly, exogenous administration of high doses of CORTs at the time of stress was shown to facilitate fear extinction and to attenuate post-traumatic stress behavior (45). In the same study, lower doses of exogenous CORT were ineffective in attenuating stress-induced abnormalities and significantly increased the propensity of mice to develop extreme behavioral responses. In our experience, a single injection of high dose of hydrocortisone can lead to an immediate transient decrease in CP expression of trafficking-related molecules, which rapidly re-bounds to higher than baseline levels (fig. S9). A possible explana-tion for this phenomenon might lie with the HPA axis activity being subjected to negative feedback control by glucocorticoids. Accordingly, acute high levels of exogenous glucocorticoid administration could lead to the suppression of both basal and stress-induced HPA axis activities.

Exposure to a single stressor (Trauma alone) did not lead to the suppression of the CP and even led to a slight trend toward an ele-vation of gene expression of trafficking-related molecules (Fig. 1B, ccl2). These results are in line with previous studies by our group (23), in which it was shown that a single short episode of stress could induce, rather than suppress, trafficking of T lymphocytes into the CNS territory. Note that while previous studies from our team reported a sponta-neous increase in trafficking molecules at the CP under mild stress (23), rather than the shutdown reported here, this increase was shown to be beneficial for coping with the stress, which is consistent with our present findings. Moreover, coping with stress was further improved by augmenting immune cell trafficking at the CP using myelin

oligodendrocyte glycoprotein immunization (23). Together, these findings might suggest that suppression of trafficking through the CP occurs only beyond a certain “threshold,” which is reached upon extremely severe or repeated episodes of stress. Thus, while different levels of stress might have different outcomes on CP function, in either case, augmenting CP gateway activity for leukocyte trafficking had a beneficial effect on stress resilience.

Augmenting CP gateway activity was previously associated with the recruitment of immunoregulatory monocytes and T cells that homed to the sites of brain pathology (19, 20). In this study, we found that increasing CP expression of leukocyte trafficking molecules was associated with an increase in leukocyte abundance in the CSF and the brain parenchyma. Among the cells detected, a dominant population was positive for the chemokine receptor CXCR3, which coincided with increased expression levels of its ligand, cxcl10, by the CP. Similarly, CXCR3+ memory T cells were previously reported to be the primary population of T cells in the CSF under physiological conditions and were suggested to infiltrate through the CP (36). In parallel to the increased number of T cells within the CNS, both systemic and local blockage of GR signaling resulted in a clear re-duction of anxiety-related behavior. The lower anxiety levels cor-related with an increase in the number of T cells recruited to the brain parenchyma, supporting the contention that these cells play a beneficial role in buffering stress responses (46). Notably, while the knockdown of CP GR alone increased the expression of trafficking- related molecules, no recruitment of cells into the brain parenchyma was detected in the absence of stress. This finding may suggest that the expression of trafficking-related molecules by the CP reflects the availability of CSF-borne immunosurveillance of the brain but will not by itself lead to unnecessary infiltration of leukocytes into the brain parenchyma.

Systemic T cells were previously suggested to support brain func-tional plasticity, including higher cognitive performance (7), neuro-genesis (5, 47), and social behavior (48), via the production of cytokines and neurotrophic factors from the CNS outer borders. Moreover, the presence of T cells at the CP compartment was proposed to support the physiological process of coping with mild stress (23). Neverthe-less, the role of T cells in the parenchyma under stress- induced anxiety-like behaviors has not been fully elucidated. Different subsets of peripheral immune cell populations have been linked to different outcomes in experimental paradigms of psychological stress. Thus, for example, monocytes were found to exacerbate neu-roinflammation and anxiety-related behavior (49, 50), whereas T cells were suggested to mediate stress resilience (10, 11). Therefore, it is possible, as was previously demonstrated both in acute CNS injury and in chronic neurodegenerative diseases (20, 28), that T cell recruitment is an interim phase in T cell–mediated immunoregu-lation of the brain microenvironment, which culminates in either local conversion of effector T cells into Tregs or further recruitment of immunosuppressive cells to the brain parenchyma and mitigation of neuroinflammatory responses (20). Single-cell RNA-seq of the T cells that homed to the CNS did not detect helper T cells expressing the transcription factors associated with the TH1 (Tbx21) or TH17 (Rorc) lineages, suggesting that the nature of the recruited cells is non-encephalitogenic, with potential anti-inflammatory activity. Specifically, TH2 cells were previously suggested to support the intrinsic anti-inflammatory properties of the brain (51), and interleukin-4 (IL-4), the central cytokine secreted under the TH2 response, was previously suggested to play a vital role in the regulation

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 11: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

11 of 15

of brain maintenance and repair (52, 53). The finding that the rel-ative ratio between the different meta-cell populations was not af-fected by reducing GR expression at the CP, and only the total number of infiltrating T cells was increased, reinforces our contention that augmenting CP gateway activity enables the boosting of a natural process that is insufficient when the stress is severe (46). Notably, our results do not imply that CORT- mediated repression of CP gateway activity restricts the migration of specific subsets of T cells; rather, we observed that enhanced gateway activity in the con-text of stress was accompanied by recruitment of GATA3 and FoxP3, but not TBX21 or RORgT-expressing T cells, to the brain pa-renchyma. Note that our findings do not exclude the potential involvement of other “non-T” immune cell populations in the re-sponse to psychological stress. While TCR+ cells were the only cell population whose abundance we showed to significantly change, it is possible that smaller immune cell populations could have been overlooked.

Overall, the present study attributes a novel negative role to CORT signaling in coping with severe psychological stress and functionally links it to limited CP-mediated immune cell trafficking to the CNS. Our results further highlight the GR expressed in CP epithelial cells as a potential target for preventing stress-induced psychopatholo-gies, including anxiety, PTSD, and depression.

MATERIALS AND METHODSAnimalsGrl1loxP/loxP mice (38) were provided by G. Schütz (German Cancer Research Center, Heidelberg, Germany). Rosa26-EYFP (40) and Foxp3.LuciDTR mice (on a C57BL/6J background) (27) were bred and maintained by the Animal Breeding Center of the Weizmann Institute of Science. C57BL/6J mice were supplied by Harlan Biotech (Jerusalem, Israel). All behavioral tests were conducted during the dark hours in a dimly lit room. All experiments were in compliance with the regulations formulated by the Institutional Animal Care and Use Committee (IACUC) of the Weizmann Institute of Science.

Stress inductionMice were exposed to two sessions of electrical foot shocks on con-secutive days, as previously described (24). On day 1, mice received 14 shocks of 1 mA, 1 s in duration over 85 min at variable intervals, representing the “trauma” in “context A.” On day 2, the same mice received five shocks of 0.7 mA, 1 s in duration over 5 min at fixed intervals, representing the “trigger” in “context B.” Shocks were ap-plied in a fear-conditioning apparatus (TSE Systems) in a transparent Plexiglas cage (21 cm by 20 cm by 36 cm) with a metal grid floor. Preexperimental habituation took place in a room adjacent to the experiment room and was illuminated with red light. Between ani-mals, the grid and cage were cleaned with 10% ethanol solution. Animals were transferred to the experiment room in darkness, and experiment room lights were kept off. Context A consisted of 10-lux illumination, without background noise. Context B consisted of 70-dB background noise and no illumination.

Systemic blockade of GRFor the systemic blockade of GR, mifepristone (RU-486; Sigma-Aldrich) was administered intraperitoneally 30 min before each foot shock session at a dose of 20 mg/kg.

RNA purification, cDNA synthesis, and real-time qPCRTotal RNA of the CP tissues, or from cell cultures, was extracted using the ZR RNA MicroPrep Kit (Zymo Research). mRNA (1 g) was converted to complementary DNA (cDNA) using the High- Capacity cDNA Reverse Transcription Kit (Applied Biosystems). The expression of specific mRNAs was assayed using fluorescence- based real-time qPCR. qPCRs were performed using Power SYBR Green PCR Master Mix (Applied Biosystems). Reactions were performed in triplicate for each sample using the CT method. Peptidylprolyl isomerase A (ppia) and hypoxanthine-guanine phos-phoribosyltransferase (hprt) were chosen as reference genes in mu-rine and human tissues, respectively, and according to their stability in the target tissue. The amplification cycles were 95°C for 5 s, 60°C for 20 s, and 72°C for 15 s. At the end of the assay, a melting curve was constructed to evaluate the specificity of the reaction. All real-time qPCRs were performed and analyzed using the StepOnePlus PCR System (Applied Biosystems).

For all murine tissues, the following primers were used: Ppia, 5′-AGCATACAGGTCCTGGCATCTTGT-3′ (forward) and 5′-CAAAGACCACATGCTTGCCATCCA-3′ (reverse); Icam1, 5′-AGATCACATTCACGGTGCTGGCTA-3′ (forward) and 5′-AGCTTTGGGATGGTAGCTGGAAGA-3′ (reverse); Ccl2, 5 ′ - C A T C C A C G T G T T G G C T C A - 3 ′ (forward) and 5′-GAT-CATCTTGCTGGTGAATGAGT-3′ (reverse); Cxcl10, 5′-AACTG-CATCCATATCGATGAC-3 ′ (forward) and 5′-GTGGCAAT-GATCTCAACAC-3′ (reverse); Gr, 5′-ACCACCTCCCAAACTCTG-3′ (forward) and 5′-GTAATTGTGCTGTCCTTCCA-3′ (reverse).

For the human ARPE cell line, the following primers were used: Hprt, 5′-CTGGCGTCGTGATTAGTG-3′ (forward) and 5′-TAAA-CACCCTTTCCAAATCCTC-3′ (reverse); Ccl2, 5′-AAAGAAGCT-GTGATCTTCAAGACC-3′ (forward) and 5′-TTCAAGTCTTCG-GAGTTTGGG-3′ (reverse); Cxcl10, 5′-GAACCTCCAGTCTCAGCA-3′ (forward) and 5′-GGTACTCCTTGAATGCCAC-3′ (reverse); Icam1, 5′-GTGACCATCTACAGCTTTCC-3′ (forward) and 5′-GCCTCA-CACTTCACTGTC-3′ (reverse).

CSF collectionCSF was collected by the cisterna magna puncture technique, as pre-viously described (7). Briefly, mice were anesthetized and placed on a stereotactic instrument so that the head formed a 135° angle with the body. A sagittal incision of the skin was made inferior to the occiput, and the subcutaneous tissue and muscle were separated. A capillary was then inserted into the cisterna magna through the dura matter lateral to the arteria dorsalis spinalis. Approximately 10 l of CSF could be aspirated from each mouse. The collected CSF was taken for analysis by flow cytometry.

Intracerebroventricular injectionsTNF (100 ng) dissolved in PBS to a final volume of 10 l was in-jected intracerebroventricularly (0.4 mm posterior to the bregma, 1.0 mm lateral to the midline, and 2.0 mm in depth from the brain surface), as described (7).

Flow cytometry sample preparation and analysisBefore all tissue collections, mice were intracardially perfused with PBS. Brains were dissected, dissociated using a gentleMACS dissoci-ator (Miltenyi Biotec), and loaded on a Percoll gradient (GE Healthcare) to isolate leukocytes. CSF samples were labeled without further pro-cessing. The following fluorochrome-labeled monoclonal antibodies

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 12: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

12 of 15

were used according to the manufacturers’ protocols: fluorescein isothiocyanate (FITC)–conjugated anti-CD45.2, FITC-conjugated anti-TCR, allophycocyanin (APC)–conjugated anti-CD11b, BV421 anti-TCR, BV421-conjugated anti-CD45, peridinin chlorophyll protein (PerCP)–Cy5.5–conjugated anti-CXCR3, and phycoerythrin (PE)-conjugated anti-CD45 (all from BioLegend) and BD Horizon v450–conjugated anti-CD4 and Alexa Fluor 700–anti-conjugated CD45.2 (BD Biosciences). Flow cytometry analysis was performed on each sample using a BD Biosciences LSRII flow cytometer, and the acquired data were analyzed using FlowJo software (Tree Star). Gating strategies for all flow cytometry analyses are shown in fig. S8.

Single-cell sortingCell populations were sorted with a SORP-Aria cytometer (BD Bio-sciences, San Jose, CA). Before sorting, all samples were filtered through a 40-m nylon mesh. Samples were gated for CD45+CD11b−TCR+ after exclusion of doublets. Isolated cells were single cell–sorted into 384-well cell capture plates containing 2 l of lysis solution and barcoded poly(T) reverse transcription (RT) primers for single-cell RNA-seq (54). Four empty wells were maintained in each 384-well plate as a no-cell control. Immediately after sorting, each plate was spun down to ensure cell immersion into the lysis solution, snap- frozen on dry ice, and stored at –80ºC until processing.

MARS-seq library preparationSingle-cell libraries were prepared as previously described (54). Briefly, mRNA from cells sorted into cell capture plates was barcoded, con-verted into cDNA, and pooled using an automated pipeline. The pooled sample was then linearly amplified by T7 in vitro transcription, and the resulting RNA was fragmented and converted into a sequencing- ready library by tagging the samples with pooled barcodes and Illumina sequences during ligation, RT, and PCR. Each pool of cells was tested for library quality, and concentration was assessed as described (54). MARS-seq libraries, pooled at equimolar concentrations, were se-quenced using an Illumina NextSeq 500 sequencer at a sequencing depth of 50,000 to 100,000 reads per cell.

Analysis of single-cell RNA-seq dataAll MARS-seq libraries were sequenced using an Illumina NextSeq 500 system at an average sequencing depth of 50,000 reads per cell. Sequences were demultiplexed, mapped, and filtered as previously described (54), extracting a set of unique molecular identifiers per cell. Cells were then clustered using the MetaCell analysis package (55). Briefly, informative genes were used to compute cell-to-cell similarity and to build a K-nn graph (k = 50) to group cells into co-hesive groups (or meta-cells). Last, the package used bootstrapping to derive strongly separated clusters. The MetaCell package is described in detail by Giladi et al. (55).

Conditional depletion of TregsDTx (8 ng g−1 body weight; Sigma-Aldrich) was injected intraperitoneally daily for four consecutive days to Foxp3.LuciDTR mice. The effi-ciency of DTx was confirmed by flow cytometry analysis of immune cells in the spleen, achieving almost complete depletion of the green fluorescent protein (GFP)–expressing FoxP3+ CD4+ Tregs (fig. S2).

Immunohistochemistry and immunocytochemistryFor staining of brain sections, two different tissue preparation pro-tocols (paraffin-embedded or microtomed free-floating sections) were

applied, as previously described (19). The primary antibodies used were rabbit anti-CD3 (1:500, Dako), rat anti-Foxp3 (1:20, eBioscience), rabbit anti-GR (1:100, Cell Signaling Technology), rat anti-ICAM (1:100, Abcam), rabbit anti-laminin (1:100, Abcam), and mouse anti– -dystroglycan (1:100, BioLegend). Secondary antibodies included Cy2/Cy3 anti-rabbit/rat/mouse antibodies (1:200, all from Jackson ImmunoResearch). The slides were exposed to Hoechst for nuclear staining (1:10,000, Invitrogen Probes) for 30 s. For Foxp3 intracel-lular staining, antigen retrieval from paraffin-embedded slides was performed using the Retrievagen Kit (BD Pharmingen). For immu-nocytochemistry, CP cells were isolated and grown on coverslips to confluence, as described before (17). Cells were washed with PBS and fixed with methanol-acetone (1:1) for 10 min at −20°C followed by two washing steps with PBS. The coverslips of the cultured CP cells were blocked with an M.O.M. immunodetection kit reagent (Vector Laboratories) containing 0.3% Triton X-100 (Sigma-Aldrich) and stained with mouse anti–ZO-1 (1:100, Invitrogen). The secondary antibody used was Cy3-conjugated donkey anti-mouse antibody (1:200, Jackson ImmunoResearch). The coverslips were exposed to Hoechst stain (1:10,000, Invitrogen) for 30 s and mounted onto slides using Immu-Mount (Thermo Scientific). For immunocytochemistry of the ARPE cell line, cells were fixed with 2.5% paraformaldehyde for 20 min, and 20% horse serum was used as a blocking agent. Mouse anti-human CD54 (Thermo Scientific) was used as primary anti-body. A fluorescence microscope (Nikon Eclipse 80i) was used for microscopic analysis. The fluorescence microscope was equipped with a digital camera (DXM 1200F, Nikon) and with 20× numerical aperture (NA) 0.50 and 40× NA 0.75 objective lenses (Plan Fluor, Nikon). Recordings were made using acquisition software (NIS- Elements, F3). To obtain an estimate of the number of labeled cells per cubic millimeter volume, the average number of cells counted in the selected sections (average surface area = 1 mm2, thickness = 0.006 mm) was multiplied by 166.66. Hoechst staining was routinely used for nuclear labeling, which served to verify quantification of the cells. Before quantification, slices were coded to mask the iden-tity of the experimental groups, and cell number was quantified by an observer blinded to the origin of the sample. To avoid overesti-mation due to counting fragments of cells that spanned several sec-tions, only cells that had an intact morphology and a nucleus that was >5 m in diameter were counted. For the quantification of cor-rected total or mean fluorescence intensity, images were analyzed using ImageJ (National Institutes of Health) via designed macros specific for each staining. Representative images were cropped, merged, and optimized using Photoshop CS6 13.0 (Adobe) and were arranged using Illustrator CS5 15.1 (Adobe).

Enzyme-linked immunosorbent assay CORT measurementsFor CORT measurements, animals were quickly anesthetized using isoflurane, and their retro-orbital blood was immediately collected into MiniCollect Z serum separator tubes (Greiner Bio-One). Serum was isolated according to the manufacturer’s protocol and snap-frozen at −80ºC until further processing. Serum CORT levels were measured using a Corticosterone ELISA kit (Enzo).

Startle responseStartle response (TSE Systems) measurements were performed as previously described (24). Briefly, mice were placed in a small Plexiglas and wire mesh cage on top of a vibration-sensitive platform in a sound- attenuated, ventilated chamber. A high-precision sensor, integrated

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 13: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

13 of 15

into the measuring platform, detected movement. Two high-frequency loudspeakers inside the chamber produced all the audio stimuli. The acoustic startle response session began with 5-min acclimation to white background noise [65 db (A)] maintained throughout the session. Two sets of 12 startle stimuli [120 db (A), 40 ms in dura-tion, with randomly varying intervals of 12 to 30 ms] were presented, interspaced by 48 startle stimuli randomly preceded by prepulses of [65 db (A) (40 ms)], [74 db (A) (40 ms)], [78 db (A) (40 ms)], or [82 db (A) (40 ms)]. Together, 72 stimuli were measured for each mouse. Latency to peak startle amplitude was measured both in response to startle stimuli and in response to startle stimuli preceded by prepulses.

Dark/Light transfer testThe apparatus for the Dark/Light transfer test (TSE Systems) con-sists of a Plexiglas box divided by a partition into two areas: one dark (14 cm by 27 cm by 26 cm) and the second brightly illuminated (30 cm by 27 cm by 26 cm, 700 lux). These areas are connected by a sliding door located at the floor level in the center of the partition. Mice were placed in the dark area, and the connecting door was opened to initiate a 5-min test session. The animal’s movements were recorded and scored using a camera and automated software (Ethovision, Noldus). Time spent in the lit arena, number of visits to the lit arena, and total distance traveled in the lit arena were mea-sured. In addition, risk assessment behavior was assessed. Percentage risk assessment time was calculated as the amount of time spent in the risk assessment arena (a rectangular zone, 3 cm by 6 cm adja-cent to the passage into the lit section) as a percentage of total time spent in the lit arena outside of the risk assessment zone.

Elevated plus mazeThe elevated plus maze apparatus consisted of a gray polyvinyl chloride maze, comprising a central part (5 cm by 5 cm), two opposing uncovered arms (30.5 cm by 5 cm), and two opposing covered arms (30.5 cm by 5 cm by 15 cm). The apparatus was elevated at a height of 53.5 cm, and the open arms were illuminated with 10 lux. Mice were placed in the center, facing an open arm to initiate a 5-min session test. The time spent and the number of entries to the uncovered arms were measured.

Open-field testMouse exploratory behavior was recorded and analyzed over a 5-min period using an automated video tracking system (VideoMot 2, V5.76) in an open-field box made of gray plastic with 50 cm by 50 cm surface area and 30-cm-high walls (TSE Systems). Tracing paths of the mice were recorded, and time spent in the corners versus the middle of the arena was evaluated.

Primary culture of CP cellsCell cultures were produced as previously described (18). Briefly, CP tissues were dissected following perfusion, digested with 0.25% trypsin (Sigma), and manually dissociated into a single-cell suspen-sion. Cells were cultured in Dulbecco’s modified Eagle’s medium (DMEM)/Ham’s F12 (Invitrogen) supplemented with 10% fetal calf serum (FCS), 1 mM l-glutamine, 1 mM sodium pyruvate, penicillin (100 U/ml), streptomycin (100 mg/ml), insulin (5 g/ml), 20 M arabinosylcytosine, sodium selenite (5 ng/ml), and epidermal growth factor (EGF) (10 ng/ml) (Sigma-Aldrich). The medium was refreshed every 24 hours, and after 72 hours, TNF (100 ng/ml) (PeproTech) was

added to the cells with or without hydrocortisone (100 ng/ml) (SOLU- CORTEF, Pfizer).

Human ARPE cell lineThe ARPE-19 cell line was cultured in 24-well plates at 1 × 105 cells per well in DMEM/Ham’s F12 medium (Invitrogen) supplemented with 10% FCS penicillin (100 U/ml) and streptomycin (100 mg/ml). Once cells reached confluence, they were treated with TNF (100 ng/ml) (PeproTech) with or without hydrocortisone (100 ng/ml) (SOLU- CORTEF, Pfizer).

TEER measurementFor the TEER measurements, CP cells were isolated as above and plated on 0.4–m–pore size, 6.5-mm-diameter polycarbonate filters (Transwell, Sigma) and analyzed in triplicate by impedance spec-troscopy using the cellZscope (nanoAnalytics, Muenster, Germany). TEER was monitored until it reached a plateau at day 5, and then both the upper and lower chambers of the Transwells were filled with medium containing TNF (100 ng/ml) and/or hydrocortisone (100 ng/ml), and TEER was monitored for an additional 72 hours.

CP-specific knockdown of GRTAT-CRE recombinase (Millipore) (40 g per mouse, 20 g in each ventricle) or PBS was injected into each of the lateral ventricles, similar to a method described elsewhere (39). Briefly, for each ven-tricle, a 10-l Hamilton syringe was used to deliver a volume of 2.11 l 0.4 mm posterior to the bregma, 1.0 mm lateral to the midline, and 2.0 mm in depth from the brain surface of anesthetized mice. Injec-tion was carried out at a rate of 1 l/min, after which the needle re-mained inside the brain for additional 4 min before being slowly removed. Mice were given a period of 2 weeks to recover from the procedure before further manipulations.

Statistical analysisThe specific tests used to analyze each set of experiments are indi-cated in the figure legends. Data were analyzed using a two-tailed Student’s t test to compare between two groups, and one-way analysis of variance (ANOVA) was used to compare several groups, followed by the Newman-Keuls post hoc procedure for pairwise comparison of groups after the null hypothesis was rejected. Sample sizes were chosen with adequate statistical power based on the literature and past experience, and mice were allocated to experimental groups according to age and genotype. Investigators were blinded to the identity of the groups during experiments and outcome assessment. All inclusion and exclusion criteria were preestablished according to the IACUC guidelines. Results are presented as mean ± SEM. In the graphs, y-axis error bars represent SEM. Statistical calculations were performed using GraphPad Prism software (GraphPad Software, San Diego, CA).

SUPPLEMENTARY MATERIALSSupplementary material for this article is available at http://advances.sciencemag.org/cgi/content/full/5/5/eaav4111/DC1Fig. S1. Decreased numbers of CD4 T cells in the brain following stress.Fig. S2. Systemic depletion of FoxP3+ Tregs.Fig. S3. FoxP3+ T cells in the HC of Treg-depleted stressed mice.Fig. S4. Augmenting CP gateway activity in the context of stress does not affect BBB integrity.Fig. S5. Infiltration of CD3+ cells to the HC in stressed mice treated with GR inhibitor.Fig. S6. Exploratory behavior in an open-field arena in stressed mice treated with GR inhibitor.Fig. S7. The effect of CORTs on the expression of trafficking-related molecules by human retinal pigment epithelial cell line.

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 14: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

14 of 15

Fig. S8. Gating strategy for flow cytometry analyses.Fig. S9. Effect of acute CORT administration on CP gene expression.Fig. S10. Meta-cell tSNE plot.

REFERENCES AND NOTES 1. B. S. McEwen, N. P. Bowles, J. D. Gray, M. N. Hill, R. G. Hunter, I. N. Karatsoreos, C. Nasca,

Mechanisms of stress in the brain. Nat. Neurosci. 18, 1353–1363 (2015). 2. S. M. Smith, W. W. Vale, The role of the hypothalamic-pituitary-adrenal axis in

neuroendocrine responses to stress. Dialogues Clin. Neurosci. 8, 383–395 (2006). 3. T. Rhen, J. A. Cidlowski, Anti-inflammatory action of glucocorticoids—New mechanisms

for old drugs. N. Engl. J. Med. 353, 1711–1723 (2005). 4. F. S. Dhabhar, Enhancing versus suppressive effects of stress on immune function:

Implications for immunoprotection and immunopathology. Neuroimmunomodulation 16, 300–317 (2009).

5. Y. Ziv, N. Ron, O. Butovsky, G. Landa, E. Sudai, N. Greenberg, H. Cohen, J. Kipnis, M. Schwartz, Immune cells contribute to the maintenance of neurogenesis and spatial learning abilities in adulthood. Nat. Neurosci. 9, 268–275 (2006).

6. A. H. Miller, C. L. Raison, The role of inflammation in depression: From evolutionary imperative to modern treatment target. Nat. Rev. Immunol. 16, 22–34 (2016).

7. K. Baruch, A. Deczkowska, E. David, J. M. Castellano, O. Miller, A. Kertser, T. Berkutzki, Z. Barnett-Itzhaki, D. Bezalel, T. Wyss-Coray, I. Amit, M. Schwartz, Aging-induced type I interferon response at the choroid plexus negatively affects brain function. Science 346, 89–93 (2014).

8. G. M. Lewitus, M. Schwartz, Behavioral immunization: Immunity to self-antigens contributes to psychological stress resilience. Mol. Psychiatry 14, 532–536 (2009).

9. G. M. Lewitus, A. Wilf-Yarkoni, Y. Ziv, M. Shabat-Simon, R. Gersner, A. Zangen, M. Schwartz, Vaccination as a novel approach for treating depressive behavior. Biol. Psychiatry 65, 283–288 (2009).

10. H. Cohen, Y. Ziv, M. Cardon, Z. Kaplan, M. A. Matar, Y. Gidron, M. Schwartz, J. Kipnis, Maladaptation to mental stress mitigated by the adaptive immune system via depletion of naturally occurring regulatory CD4+CD25+ cells. J. Neurobiol. 66, 552–563 (2006).

11. R. B. Scheinert, M. H. Haeri, M. L. Lehmann, M. Herkenham, Therapeutic effects of stress-programmed lymphocytes transferred to chronically stressed mice. Prog. Neuropsychopharmacol. Biol. Psychiatry 70, 1–7 (2016).

12. A. Louveau, T. H. Harris, J. Kipnis, Revisiting the mechanisms of CNS immune privilege. Trends Immunol. 36, 569–577 (2015).

13. M. Schwartz, K. Baruch, The resolution of neuroinflammation in neurodegeneration: Leukocyte recruitment via the choroid plexus. EMBO J. 33, 7–22 (2014).

14. G. Moalem, R. Leibowitz–Amit, E. Yoles, F. Mor, I. R. Cohen, M. Schwartz, Autoimmune T cells protect neurons from secondary degeneration after central nervous system axotomy. Nat. Med. 5, 49–55 (1999).

15. O. Rapalino, O. Lazarov-Spiegler, E. Agranov, G. J. Velan, E. Yoles, M. Fraidakis, A. Soloman, R. Gepstein, A. Katz, M. Belkin, M. Hadani, M. Schwartz, Implantation of stimulated homologous macrophages results in partial recovery of paraplegic rats. Nat. Med. 4, 814–821 (1998).

16. R. Shechter, O. Miller, G. Yovel, N. Rosenzweig, A. London, J. Ruckh, K. W. Kim, E. Klein, V. Kalchenko, P. Bendel, S. A. Lira, S. Jung, M. Schwartz, Recruitment of beneficial M2 macrophages to injured spinal cord is orchestrated by remote brain choroid plexus. Immunity 38, 555–569 (2013).

17. G. Kunis, K. Baruch, N. Rosenzweig, A. Kertser, O. Miller, T. Berkutzki, M. Schwartz, IFN-–dependent activation of the brain’s choroid plexus for CNS immune surveillance and repair. Brain 136, 3427–3440 (2013).

18. K. Baruch, A. Kertser, Z. Porat, M. Schwartz, Cerebral nitric oxide represses choroid plexus NFB-dependent gateway activity for leukocyte trafficking. EMBO J. 34, 1816–1828 (2015).

19. K. Baruch, N. Rosenzweig, A. Kertser, A. Deczkowska, A. M. Sharif, A. Spinrad, A. Tsitsou-Kampeli, A. Sarel, L. Cahalon, M. Schwartz, Breaking immune tolerance by targeting Foxp3+ regulatory T cells mitigates Alzheimer’s disease pathology. Nat. Commun. 6, 7967 (2015).

20. G. Kunis, K. Baruch, O. Miller, M. Schwartz, Immunization with a myelin-derived antigen activates the brain’s choroid plexus for recruitment of immunoregulatory cells to the CNS and attenuates disease progression in a mouse model of ALS. J. Neurosci. 35, 6381–6393 (2015).

21. S. Dá Mesquita, A. C. Ferreira, F. Gao, G. Coppola, D. H. Geschwind, J. C. Sousa, M. Correia-Neves, N. Sousa, J. A. Palha, F. Marques, The choroid plexus transcriptome reveals changes in type I and II interferon responses in a mouse model of Alzheimer’s disease. Brain Behav. Immun. 49, 280–292 (2015).

22. K. Baruch, N. Ron-Harel, H. Gal, A. Deczkowska, E. Shifrut, W. Ndifon, N. Mirlas-Neisberg, M. Cardon, I. Vaknin, L. Cahalon, T. Berkutzki, M. P. Mattson, F. Gomez-Pinilla, N. Friedman, M. Schwartz, CNS-specific immunity at the choroid plexus shifts toward destructive TH2 inflammation in brain aging. Proc. Natl. Acad. Sci. U.S.A. 110, 2264–2269 (2013).

23. G. M. Lewitus, H. Cohen, M. Schwartz, Reducing post-traumatic anxiety by immunization. Brain Behav. Immun. 22, 1108–1114 (2008).

24. M. Lebow, A. Neufeld-Cohen, Y. Kuperman, M. Tsoory, S. Gil, A. Chen, Susceptibility to PTSD-like behavior is mediated by corticotropin-releasing factor receptor type 2 levels in the bed nucleus of the stria terminalis. J. Neurosci. 32, 6906–6916 (2012).

25. J. C. Porter, A. Hall, Epithelial ICAM-1 and ICAM-2 regulate the egression of human T cells across the bronchial epithelium. FASEB J. 23, 492–502 (2009).

26. N. Murugesan, D. Paul, Y. Lemire, B. Shrestha, S. Ge, J. S. Pachter, Active induction of experimental autoimmune encephalomyelitis by MOG35-55 peptide immunization is associated with differential responses in separate compartments of the choroid plexus. Fluids Barriers CNS 9, 15 (2012).

27. J. Suffner, K. Hochweller, M.-C. Kühnle, X. Li, R. A. Kroczek, N. Garbi, G. J. Hämmerling, Dendritic cells support homeostatic expansion of Foxp3+ regulatory T cells in Foxp3.LuciDTR mice. J. Immunol. 184, 1810–1820 (2010).

28. C. Raposo, N. Graubardt, M. Cohen, C. Eitan, A. London, T. Berkutzki, M. Schwartz, CNS repair requires both effector and regulatory T cells with distinct temporal and spatial profiles. J. Neurosci. 34, 10141–10155 (2014).

29. M. Saito, T. Iwawaki, C. Taya, H. Yonekawa, M. Noda, Y. Inui, E. Mekada, Y. Kimata, A. Tsuru, K. Kohno, Diphtheria toxin receptor-mediated conditional and targeted cell ablation in transgenic mice. Nat. Biotechnol. 19, 746–750 (2001).

30. Y. Liu, I. Teige, B. Birnir, S. Issazadeh-Navikas, Neuron-mediated generation of regulatory T cells from encephalitogenic T cells suppresses EAE. Nat. Med. 12, 518–525 (2006).

31. S. Agrawal, P. Anderson, M. Durbeej, N. van Rooijen, F. Ivars, G. Opdenakker, L. M. Sorokin, Dystroglycan is selectively cleaved at the parenchymal basement membrane at sites of leukocyte extravasation in experimental autoimmune encephalomyelitis. J. Exp. Med. 203, 1007–1019 (2006).

32. S. N. Shirazi, A. R. Friedman, D. Kaufer, S. A. Sakhai, Glucocorticoids and the brain: Neural mechanisms regulating the stress response. Adv. Exp. Med. Biol. 872, 235–252 (2015).

33. R. B. Auwardt, S. J. Mudge, C. G. Chen, D. A. Power, Regulation of nuclear factor kappaB by corticosteroids in rat mesangial cells. J. Am. Soc. Nephrol. 9, 1620–1628 (1998).

34. H. Derendorf, H. Möllmann, J. Barth, C. Möllmann, S. Tunn, M. Krieg, Pharmacokinetics and oral bioavailability of hydrocortisone. J. Clin. Pharmacol. 31, 473–476 (1991).

35. J. P. O’Callaghan, K. A. Kelly, A. R. Locker, D. B. Miller, S. M. Lasley, Corticosterone primes the neuroinflammatory response to DFP in mice: Potential animal model of Gulf War Illness. J. Neurochem. 133, 708–721 (2015).

36. P. Kivisäkk, C. Trebst, Z. Liu, B. H. Tucky, T. L. Sørensen, R. A. Rudick, M. Mack, R. M. Ransohoff, T-cells in the cerebrospinal fluid express a similar repertoire of inflammatory chemokine receptors in the absence or presence of CNS inflammation: Implications for CNS trafficking. Clin. Exp. Immunol. 129, 510–518 (2002).

37. B. Srinivasan, A. R. Kolli, M. B. Esch, H. E. Abaci, M. L. Shuler, J. J. Hickman, TEER measurement techniques for in vitro barrier model systems. J. Lab. Autom. 20, 107–126 (2015).

38. F. Tronche, C. Kellendonk, O. Kretz, P. Gass, K. Anlag, P. C. Orban, R. Bock, R. Klein, G. Schütz, Disruption of the glucocorticoid receptor gene in the nervous system results in reduced anxiety. Nat. Genet. 23, 99–103 (1999).

39. J. Spatazza, H. H. C. Lee, A. A. di Nardo, L. Tibaldi, A. Joliot, T. K. Hensch, A. Prochiantz, Choroid-plexus–derived Otx2 homeoprotein constrains adult cortical plasticity. Cell Rep. 3, 1815–1823 (2013).

40. S. Srinivas, T. Watanabe, C. S. Lin, C. M. William, Y. Tanabe, T. M. Jessell, F. Costantini, Cre reporter strains produced by targeted insertion of EYFP and ECFP into the ROSA26 locus. BMC Dev. Biol. 1, 4 (2001).

41. F. Marques, J. C. Sousa, G. Coppola, A. M. Falcao, A. J. Rodrigues, D. H. Geschwind, N. Sousa, M. Correia-Neves, J. A. Palha, Kinetic profile of the transcriptome changes induced in the choroid plexus by peripheral inflammation. J. Cereb. Blood Flow Metab. 29, 921–932 (2009).

42. S. Balusu, E. van Wonterghem, R. de Rycke, K. Raemdonck, S. Stremersch, K. Gevaert, M. Brkic, D. Demeestere, V. Vanhooren, A. Hendrix, C. Libert, R. E. Vandenbroucke, Identification of a novel mechanism of blood-brain communication during peripheral inflammation via choroid plexus-derived extracellular vesicles. EMBO Mol. Med. 8, 1162–1183 (2016).

43. N. A. S. Rao, M. T. McCalman, P. Moulos, K.-J. Francoijs, A. Chatziioannou, F. N. Kolisis, M. N. Alexis, D. J. Mitsiou, H. G. Stunnenberg, Coactivation of GR and NFKB alters the repertoire of their binding sites and target genes. Genome Res. 21, 1404–1416 (2011).

44. D. de Quervain, L. Schwabe, B. Roozendaal, Stress, glucocorticoids and memory: Implications for treating fear-related disorders. Nat. Rev. Neurosci. 18, 7–19 (2017).

45. H. Cohen, M. A. Matar, D. Buskila, Z. Kaplan, J. Zohar, Early post-stressor intervention with high-dose corticosterone attenuates posttraumatic stress response in an animal model of posttraumatic stress disorder. Biol. Psychiatry 64, 708–717 (2008).

46. M. Schwartz, K. Baruch, Vaccine for the mind: Immunity against self at the choroid plexus for erasing biochemical consequences of stressful episodes. Hum. Vaccin. Immunother. 8, 1465–1468 (2012).

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 15: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

Kertser et al., Sci. Adv. 2019; 5 : eaav4111 29 May 2019

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

15 of 15

47. S. A. Wolf, B. Steiner, A. Wengner, M. Lipp, T. Kammertoens, G. Kempermann, Adaptive peripheral immune response increases proliferation of neural precursor cells in the adult hippocampus. FASEB J. 23, 3121–3128 (2009).

48. A. J. Filiano, Y. Xu, N. J. Tustison, R. L. Marsh, W. Baker, I. Smirnov, C. C. Overall, S. P. Gadani, S. D. Turner, Z. Weng, S. N. Peerzade, H. Chen, K. S. Lee, M. M. Scott, M. P. Beenhakker, V. Litvak, J. Kipnis, Unexpected role of interferon- in regulating neuronal connectivity and social behaviour. Nature 535, 425–429 (2016).

49. D. B. McKim, M. D. Weber, A. Niraula, C. M. Sawicki, X. Liu, B. L. Jarrett, K. Ramirez-Chan, Y. Wang, R. M. Roeth, A. D. Sucaldito, C. G. Sobol, N. Quan, J. F. Sheridan, J. P. Godbout, Microglial recruitment of IL-1–producing monocytes to brain endothelium causes stress-induced anxiety. Mol. Psychiatry 23, 1421–1431 (2018).

50. E. S. Wohleb, N. D. Powell, J. P. Godbout, J. F. Sheridan, Stress-induced recruitment of bone marrow-derived monocytes to the brain promotes anxiety-like behavior. J. Neurosci. 33, 13820–13833 (2013).

51. U. Gimsa, S. A. Wolf, D. Haas, I. Bechmann, R. Nitsch, Th2 cells support intrinsic anti-inflammatory properties of the brain. J. Neuroimmunol. 119, 73–80 (2001).

52. X. Zhao, H. Wang, G. Sun, J. Zhang, N. J. Edwards, J. Aronowski, Neuronal interleukin-4 as a modulator of microglial pathways and ischemic brain damage. J. Neurosci. 35, 11281–11291 (2015).

53. J. T. Walsh, S. Hendrix, F. Boato, I. Smirnov, J. Zheng, J. R. Lukens, S. Gadani, D. Hechler, G. Gölz, K. Rosenberger, T. Kammertöns, J. Vogt, C. Vogelaar, V. Siffrin, A. Radjavi, A. Fernandez-Castaneda, A. Gaultier, R. Gold, T. D. Kanneganti, R. Nitsch, F. Zipp, J. Kipnis, MHCII-independent CD4+ T cells protect injured CNS neurons via IL-4. J. Clin. Invest. 125, 699–714 (2015).

54. D. A. Jaitin, E. Kenigsberg, H. Keren-Shaul, N. Elefant, F. Paul, I. Zaretsky, A. Mildner, N. Cohen, S. Jung, A. Tanay, I. Amit, Massively parallel single-cell RNA-seq for marker-free decomposition of tissues into cell types. Science 343, 776–779 (2014).

55. A. Giladi, F. Paul, Y. Herzog, Y. Lubling, A. Weiner, I. Yofe, D. Jaitin, N. Cabezas-Wallscheid, R. Dress, F. Ginhoux, A. Trumpp, A. Tanay, I. Amit, Single-cell characterization of haematopoietic progenitors and their trajectories in homeostasis and perturbed haematopoiesis. Nat. Cell Biol. 20, 836–846 (2018).

Acknowledgments: We thank G. Schütz and A. Chen for providing and assisting us with GR-floxed mice, O. Matcovitch-Natan for consulting on flow cytometry sorting techniques, and S. Schwarzbaum for editing the manuscript. Funding: This study was supported by Advanced European Research Council grants (232835), the Minerva Foundation, and an ISF Legacy BioMed grant (1353/15). A.D. is supported by Steven and Eden Romick. M.S. holds the Maurice and Ilse Katz Professorial Chair in Neuroimmunology. M.T. is the incumbent of the Carolito Stiftung Research Fellow Chair in Neurodegenerative Diseases. Author contributions: A.K. and K.B., in equal contribution and under the mentoring of M.S., conceived the general ideas of this study, performed all of the experiments, analyzed the data, and prepared the data for presentation. A.D., M.K., and T.C. assisted in experimental procedures. A.W. and I.A. assisted in analyzing data. I.C. assisted in transepithelial resistance recording experiments and subsequent data analysis. M.T. helped planning and assisted with all animal models. S.B.-H. assisted in intracerebroventricular injections. A.K., K.B., and M.S. wrote the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data related to this paper may be requested from the authors.

Submitted 13 September 2018Accepted 22 April 2019Published 29 May 201910.1126/sciadv.aav4111

Citation: A. Kertser, K. Baruch, A. Deczkowska, A. Weiner, T. Croese, M. Kenigsbuch, I. Cooper, M. Tsoory, S. Ben-Hamo, I. Amit, M. Schwartz, Corticosteroid signaling at the brain-immune interface impedes coping with severe psychological stress. Sci. Adv. 5, eaav4111 (2019).

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from

Page 16: PSYCHOLOGY Copyright © 2019 Corticosteroid signaling at ......Trauma only or Trauma + Trigger. Gray area represents the SE margins of the nonstressed group (time: −24 hours, before

psychological stressCorticosteroid signaling at the brain-immune interface impedes coping with severe

M. SchwartzA. Kertser, K. Baruch, A. Deczkowska, A. Weiner, T. Croese, M. Kenigsbuch, I. Cooper, M. Tsoory, S. Ben-Hamo, I. Amit and

DOI: 10.1126/sciadv.aav4111 (5), eaav4111.5Sci Adv 

ARTICLE TOOLS http://advances.sciencemag.org/content/5/5/eaav4111

MATERIALSSUPPLEMENTARY http://advances.sciencemag.org/content/suppl/2019/05/23/5.5.eaav4111.DC1

REFERENCES

http://advances.sciencemag.org/content/5/5/eaav4111#BIBLThis article cites 55 articles, 15 of which you can access for free

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Terms of ServiceUse of this article is subject to the

is a registered trademark of AAAS.Science AdvancesYork Avenue NW, Washington, DC 20005. The title (ISSN 2375-2548) is published by the American Association for the Advancement of Science, 1200 NewScience Advances

License 4.0 (CC BY-NC).Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial Copyright © 2019 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of

on Novem

ber 17, 2020http://advances.sciencem

ag.org/D

ownloaded from