Memory Deficits Induced by Inflammation Are Regulated by &alpha

151 downloads 0 Views 2MB Size Report
Sep 27, 2012 - ... London, Ontario, N6A 5K8, Canada. 6Department of Anesthesia, Sunnybrook Health Sciences Centre, Toronto, Ontario, M4N 3M5, Canada.
Cell Reports

Report Memory Deficits Induced by Inflammation Are Regulated by a5-Subunit-Containing GABAA Receptors Dian-Shi Wang,1 Agnieszka A. Zurek,1 Irene Lecker,1 Jieying Yu,1 Armen M. Abramian,3 Sinziana Avramescu,2 Paul A. Davies,3 Stephen J. Moss,3,4 Wei-Yang Lu,5 and Beverley A. Orser1,2,6,* 1Department

of Physiology of Anesthesia University of Toronto, Toronto, Ontario, M5S 1A8, Canada 3Department of Neuroscience, Tufts University, Boston, MA 02111, USA 4Department of Neuroscience, Physiology and Pharmacology, University College, London, WC1E 6BT, UK 5Department of Physiology and Pharmacology, University of Western Ontario, London, Ontario, N6A 5K8, Canada 6Department of Anesthesia, Sunnybrook Health Sciences Centre, Toronto, Ontario, M4N 3M5, Canada *Correspondence: [email protected] http://dx.doi.org/10.1016/j.celrep.2012.08.022 2Department

SUMMARY

Systemic inflammation causes learning and memory deficits through mechanisms that remain poorly understood. Here, we studied the pathogenesis of memory loss associated with inflammation and found that we could reverse memory deficits by pharmacologically inhibiting a5-subunit-containing g-aminobutyric acid type A (a5GABAA) receptors and deleting the gene associated with the a5 subunit. Acute inflammation reduces long-term potentiation, a synaptic correlate of memory, in hippocampal slices from wild-type mice, and this reduction was reversed by inhibition of a5GABAA receptor function. A tonic inhibitory current generated by a5GABAA receptors in hippocampal neurons was increased by the key proinflammatory cytokine interleukin-1b through a p38 mitogen-activated protein kinase signaling pathway. Interleukin-1b also increased the surface expression of a5GABAA receptors in the hippocampus. Collectively, these results show that a5GABAA receptor activity increases during inflammation and that this increase is critical for inflammation-induced memory deficits. INTRODUCTION Acute systemic inflammation caused by a variety of disorders, including autoimmune diseases, infection, traumatic brain injury, and stroke, can lead to memory loss (Dantzer et al., 2008; Di Filippo et al., 2008; Kipnis et al., 2008; Yirmiya and Goshen, 2011). Such memory loss manifests as impaired explicit recall in humans, and deficiencies of fear-associated memory and reduced performance for object-recognition tasks in laboratory animals (Yirmiya and Goshen, 2011). Inflammation also contributes to chronic neurodegenerative diseases that are characterized by

memory loss, including Alzheimer disease, Parkinson disease, multiple sclerosis, and even AIDS-related dementia (Di Filippo et al., 2008; Kipnis et al., 2008; Yirmiya and Goshen, 2011). Systemic inflammation increases the production of multiple cytokines in the brain, including interleukin-1b (IL-1b), tumor necrosis factor-a (TNF-a), and IL-6 (Dantzer et al., 2008; Yirmiya and Goshen, 2011). In particular, elevated levels of IL-1b are known to contribute to inflammation-induced memory deficits. For example, in patients with sepsis-associated encephalopathy, increased plasma levels of IL-1b were correlated with cognitive deficits (Serantes et al., 2006). In another study, elderly people with a genetic variant of the IL-1b-converting enzyme that produces lower levels of IL-1b exhibited better cognitive performance than the general aged population (Trompet et al., 2008). In laboratory animals that underwent orthopedic surgery, elevated levels of IL-1b in the hippocampus were strongly correlated with memory deficits (Cibelli et al., 2010). Furthermore, it was shown that memory deficits associated with elevated levels of IL-1b were typically hippocampus dependent, whereas hippocampus-independent memory was spared (Yirmiya and Goshen, 2011). The mechanisms underlying inflammation-induced memory deficits are not well understood. Changes involving multiple neurotransmitter receptors have been demonstrated, including a reduction in N-methyl-D-aspartate (NMDA) receptor activity, alterations in the trafficking and phosphorylation of 2-amino-3(5-methyl-3-oxo-1,2-oxazol-4-yl)propanoic acid receptors, and activation of the a7 subtype of nicotinic acetylcholine receptors (Stellwagen et al., 2005; Terrando et al., 2011; Viviani et al., 2003). However, despite a long history of investigation, no treatments that effectively reverse or prevent the memory deficits associated with inflammation are available. General inhibition of the inflammatory response and specific blocking of IL-1b activity by inhibiting the membrane-bound type 1 IL-1 receptors are impractical approaches because of the risks of infection and delayed wound healing (Fleischmann et al., 2006). Moreover, low basal levels of IL-1b in the hippocampus play a physiological role in maintaining normal memory performance (Yirmiya

488 Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors

and Goshen, 2011). Thus, identification of additional downstream mediators of inflammation that induce memory loss is necessary for the development of effective treatments. The inhibitory neurotransmitter g-aminobutyric acid (GABA) is a powerful regulator of learning, memory, and synaptic plasticity (Collinson et al., 2002). GABA type A (GABAA) receptors generate two distinct forms of inhibition: phasic, fast inhibitory postsynaptic currents and a tonic form of inhibition that is primarily mediated by extrasynaptic GABAA receptors (Brickley and Mody, 2012; Glykys and Mody, 2007; Luscher et al., 2011). In the CA1 subfield of the hippocampus, a tonic inhibitory conductance is generated mainly by the a5-subunit-containing subtype of GABAA (a5GABAA) receptors (Caraiscos et al., 2004), which are likely composed of a5b3g2 subunits (Burgard et al., 1996; Glykys and Mody, 2007; Ju et al., 2009). Drugs that increase a5GABAA receptor activity cause profound memory blockade (Cheng et al., 2006; Martin et al., 2009). Conversely, a reduction in the expression or function of this receptor improves performance for certain memory tasks (Collinson et al., 2002; Martin et al., 2010). Here, using a combination of behavioral studies, electrophysiological recordings, and biochemical methods together with genetic and pharmacological tools, we sought to determine whether increased a5GABAA receptor activity contributes to inflammation-induced memory deficits. The results of these studies show that a5GABAA receptors are critical downstream mediators of inflammation-induced memory deficits. RESULTS a5GABAA Receptors Regulate Inflammation-Induced Impairment of Contextual Fear Memory We first investigated the role of a5GABAA receptors in inflammation-induced deficits using a fear-associated memory paradigm (Fanselow and Poulos, 2005). Wild-type (WT) and a5-subunit null mutant (Gabra5 / ) mice were treated with IL-1b (1 mg kg 1, i.p.) to mimic acute systemic inflammation (Yirmiya and Goshen, 2011). Three hours after treatment, the mice were trained to associate an electric foot shock (unconditioned stimulus) with an environmental context and a tone (conditioned stimuli). To study contextual fear memory, which is hippocampus dependent (Fanselow and Poulos, 2005), we reexposed the mice to the same environmental context 24 hr after training. To study cued fear memory in response to tone, which is hippocampus independent (Fanselow and Poulos, 2005), we reexposed the mice to the tone after training. The percentage of time that each mouse spent freezing in the same context or in response to the tone was used to assess memory. IL-1b-treated WT mice showed impairment of contextual fear memory, as evidenced by lower freezing scores compared with vehicle-treated controls (Figure 1A; n = 14–16; two-way analysis of variance [ANOVA], effect of IL-1b, F(1,56) = 5.60, p = 0.02; effect of genotype, F(1,56) = 0.07, p = 0.79; effect of interaction, F(1,56) = 2.86, p = 0.10; Bonferroni post hoc test, p < 0.05), as shown previously (Yirmiya and Goshen, 2011). In contrast, Gabra5 / mice treated with IL-1b exhibited no memory deficits (Figure 1A; n = 14–16; Bonferroni post hoc test, p > 0.05). To determine whether pharmacological inhibition of a5GABAA receptors would prevent the memory deficit, we studied the

effects of two benzodiazepine inverse agonists that preferentially inhibit a5GABAA receptors (Atack et al., 2009; Quirk et al., 1996). At 30 min before fear conditioning, the mice were treated with either L-655,708 [ethyl (13aS)-7-methoxy9-oxo-11,12,13,13a-tetrahydro-9H-imidazo[1,5-a]pyrrolo[2,1-c] [1,4]benzodiazepine-1-carboxylate] (0.35 or 0.5 mg kg 1, i.p.) or MRK-016 [3-tert-butyl-7-(5-methylisoxazol-3-yl)-2-(1-methyl1H-1,2,4-triazol-5-ylmethoxy)-pyrazolo[1,5-d][1,2,4]triazine] (3 mg kg 1, i.p.). Of note, under baseline conditions, L-655,708 does not modify contextual fear memory in WT or Gabra5 / mice (Martin et al., 2010). However, L-655,708 reversed the contextual fear memory deficits induced by IL-1b in WT mice (Figure 1B; n = 11; Student’s t test, t = 2.0, p = 0.03). Similarly, another inverse agonist that is structurally distinct from L-655,708, MRK016, attenuated the contextual fear memory deficits induced by IL-1b in WT mice (Figure 1B; n = 16; Student’s t test, t = 2.1, p = 0.04). We next used a widely employed model of systemic inflammation to probe whether elevated levels of endogenous cytokines also impair memory through activation of a5GABAA receptors. WT and Gabra5 / mice were treated with the gram-negative bacterial endotoxin lipopolysaccharide (LPS; 125 mg kg 1, i.p.) 3 hr before behavioral training. WT mice, but not Gabra5 / mice, exhibited a reduction in contextual fear memory after treatment with LPS (Figure 1C; n = 10–15; two-way ANOVA, effect of LPS, F(1,53) = 8.16, p = 0.006; effect of genotype, F(1,53) = 0.0005, p = 0.98; effect of interaction, F(1,53) = 1.02, p = 0.32; Bonferroni post hoc test, p < 0.05 for WT). The deficit exhibited by WT mice was reversed by inhibiting a5GABAA receptors with L-655,708 (Figure 1C; n = 10–11; Student’s t test, t = 2.7, p = 0.01). Of interest, treatment with either IL-1b or LPS did not affect cued fear memory in response to the conditioned tone (Figures 1D and S1; n = 14–16; two-way ANOVA, effect of IL-1b, F(1,54) = 1.28, p = 0.26; effect of genotype, F(1,54) = 1.45, p = 0.23; effect of interaction, F(1,54) = 0.06, p = 0.80). Collectively, these results suggest that the impairment of memory induced by inflammation did not result from a global disruption of cognitive processes, but rather from disruption of hippocampal function. a5GABAA Receptors Regulate Inflammation-Induced Impairment of Long-Term Potentiation Long-term potentiation (LTP), which refers to the prolonged enhancement of excitatory synaptic transmission, is a widely studied cellular model of learning and memory (Lynch, 2004). We next examined whether the activity of a5GABAA receptors regulates the impairment of LTP induced by acute inflammation. Brain slices were prepared from mice 3 hr after they were injected with the endotoxin LPS (125 mg kg 1, i.p.). Theta burst stimulation of the Schaffer collateral pathway was used to induce LTP in the CA1 subfield of the hippocampus, and field postsynaptic potentials (fPSPs) were recorded before and after stimulation (Figure 2A, inset). In slices from the vehicle-treated WT mice, after stimulation, the slope of the fPSPs increased to 136.1% ± 5.6% of baseline (n = 9). In contrast, in slices from the LPStreated mice, the increase in LTP was only 113.1% ± 2.5% of baseline (n = 10; one-way ANOVA, F(2,26) = 4.38, p = 0.02; Newman-Keuls post hoc test, p < 0.05; Figures 2A and 2B).

Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors 489

A

B

C

D

Figure 1. Inflammation-Induced Impairment of Contextual Fear Memory Is Absent in Gabra5 / cological Inhibition of a5GABAA Receptor in WT Mice

Mice and Can Be Prevented by Pharma-

(A) IL-1b reduced the freezing scores for contextual fear memory only in WT mice. The bar graph shows the summarized results. (B) L-655,708 or MRK-016 restored the freezing scores for contextual fear memory to control values in WT mice. (C) LPS reduced the freezing scores for contextual fear memory only in WT mice, and L-655,708 restored the freezing scores to control values in WT mice. (D) IL-1b did not affect the freezing scores for cued fear memory in response to tone in both WT and Gabra5 / mice. Data are represented as mean ± SEM. Here and in subsequent figures, Con stands for control. *p < 0.05, **p < 0.01, ***p < 0.001, N.S.: nonsignificant result. See also Figure S1.

490 Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors

fPSP slope (% baseline) fPSP slope (% baseline)

200

Con LPS

150 100

Stimulation

50 Stimulation

0

0

10

20

Recording

CA1 CA3

30 40 50 Time (min)

60

70

Gabra5-/-

200

Con LPS

150 100

IL-1b Increases the Tonic Current in the CA1 region In the CA1 region of the hippocampus, a tonic inhibitory conductance generated by a5GABAA receptors regulates synaptic plasticity and memory processes (Collinson et al., 2002; Martin et al., 2010), and an increase in the activity of these receptors causes memory deficits (Cheng et al., 2006; Martin et al., 2009). Therefore, we next determined whether IL-1b increases the amplitude of the tonic inhibitory current in CA1 pyramidal neurons by recording whole-cell currents in brain slices from WT mice injected with IL-1b (1 mg kg 1, i.p., 2–3 hr before sacrifice) or vehicle control (0.1% bovine serum albumin in phosphatebuffered saline). To investigate the changes in tonic current, the GABAA receptor antagonist bicuculline (Bic; 10 mM) or L-655,708 (20 nM) was applied, and the reduction in holding current was measured (Caraiscos et al., 2004; Glykys and Mody, 2007). In slices from IL-1b-treated mice, a 2-fold increase in tonic current density revealed by either Bic or L-655,708 was observed relative to vehicle-treated controls (Student’s t test, n = 11–14, t = 2.2, p = 0.04 for Bic; n = 6–10, t = 2.3, p = 0.04 for L-655,708; Figure 3A).

50 Stimulation

0

0

10

20

B fPSP slope (% baseline)

post hoc test, p < 0.05; Figure 2B). Furthermore, in slices from Gabra5 / mice, LPS failed to impair LTP. The fPSP slope was 133.1% ± 4.3% of baseline for LPS treatment (n = 15) compared with 135.4% ± 5.9% of baseline for vehicle treatment (n = 13; Figures 2A and 2B), and there was no significant difference compared with the WT (two-way ANOVA, effect of LPS, F(1,43) = 6.33, p = 0.02; effect of genotype, F(1,43) = 3.71, p = 0.06; effect of interaction, F(1,43) = 4.28, p = 0.04; Bonferroni post hoc test, p > 0.05 for Gabra5 / ). Together, these results suggest that a5GABAA receptors are required for inflammation-induced impairment of LTP. Next, to determine whether the LPS-induced reduction in LTP was mediated in part by IL-1b binding to the IL-1 receptor, as was suggested previously (Lynch et al., 2004), slices from LPStreated mice were incubated with the specific IL-1 receptor antagonist (IL-1ra; 100 ng ml 1 for 1 hr before recording). IL1ra restored LTP to 126.2% ± 3.9% of baseline (n = 10; p < 0.05 compared with LPS; Figure S2), which confirmed that the reduction in LTP was largely mediated by an increase in IL-1b activity.

WT

A

30 40 50 Time (min)

60

70

N.S. 150 100 50 0

WT

Gabra5-/-

Figure 2. Inflammation Attenuates LTP in CA1 Region Only in WT Mice and This Can Be Prevented by Pharmacological Inhibition of a5GABAA Receptor (A) LPS impaired LTP induced by theta burst stimulation only in slices from WT mice. See also Figure S2. (B) Summarized data showing the last 5 min of the recording. Note that pharmacological inhibition of a5GABAA receptors with L-655,708 prevented the impairment of LTP by LPS. Perfusion of the slices with L-655,708 (20 nM) was started 10 min before the stimulation and was present for LTP recording. Data are represented as mean ± SEM.

Pharmacological inhibition of a5GABAA receptors in hippocampal slices by application of L-655,708 (20 nM) eliminated LPS-induced reduction of LTP (fPSP slope: 130.3% ± 7.7% of baseline, n = 10; one-way ANOVA followed by Newman-Keuls

IL-1b Increases a5GABAA Receptor Activity in Cultured Neurons To investigate whether IL-1b directly enhances the tonic inhibitory conductance, whole-cell currents were recorded from cultured hippocampal neurons. The cell culture preparation offers the advantage of a high-throughput assay that can be used to study the effects of proinflammatory cytokines on the tonic and synaptic inhibitory currents. Treatment of the neurons with IL-1b (20 ng ml 1 for 20 min) increased the amplitude of the tonic current by 45% (IL-1b 1.6 ± 0.1 pA pF 1, n = 22, versus control 1.1 ± 0.1 pA pF 1, n = 21; one-way ANOVA, F(4,80) = 7.13, p < 0.0001; Dunnett’s post hoc test, p < 0.001 compared with control; Figure 3B). Increasing the concentration and duration of IL-1b treatment (60 ng ml 1 for 3 h) further increased the current amplitude (IL-1b 2.0 ± 0.3 pA pF 1, n = 6, versus control 1.2 ± 0.2 pA pF 1, n = 6; Student’s t test, t = 2.21, p = 0.03). No

Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors 491

Figure 3. IL-1b Increases the Tonic Current

A

(A) The tonic current density revealed by Bic (10 mM) and L-655,708 (20 nM) was increased 2-fold in slices from mice treated with IL-1b (1.0 mg kg 1) compared with vehicle-treated controls. (B–G) IL-1b increased the tonic current generated by a5GABAA receptors in cultured hippocampal neurons. (B) Exogenous IL-1b (20 min) increased the tonic current, as revealed by the application of Bic (100 mM, left). The concentration-dependent effects of IL-1b on tonic current density are shown on the right (n = 13–22). (C) IL-1b enhancement of tonic current was blocked by coapplication of IL-1ra (250 ng ml 1, 30 min). (D) A representative recording showing the tonic currents revealed by Bic (100 mM) or by the inverse agonist for a5GABAA receptors, L-655,708 (20 nM). (E and F) a5GABAA receptors are necessary for the enhancing effects of IL-1b on the tonic current revealed by (E) Bic (100 mM) and (F) L-655,708 (20 nM); n = 7–13; two-way ANOVA. (E) Effect of IL-1b: F(1,37) = 9.19, p = 0.004; effect of genotype: F(1,37) = 31.14, p < 0.0001; effect of interaction: F(1,37) = 4.35, p = 0.04. (F) Effect of IL-1b: F(1,36) = 8.17, p = 0.007; effect of genotype: F(1,36) = 138.92, p < 0.0001; effect of interaction: F(1,36) = 10.43, p = 0.03. Bonferroni post hoc test, **p < 0.01, ***p < 0.001 compared with control. (G) The tonic current was increased by treating neuron and microglia cocultures, but not neurons cultured alone, with LPS (100 ng ml 1, overnight). The LPS-enhancing effects on the tonic current were blocked by treating the cocultures with IL1ra (250 ng ml 1, overnight). Data are represented as mean ± SEM. See also Figure S3 and Table S1.

B

C

F

D

E

G

further increase in current amplitude was observed when neurons were treated for 12–15 hr (IL-1b 20 ng ml 1 1.3 ± 0.6 pA pF 1, n = 21, versus control 0.8 ± 0.08 pA pF 1, n = 12; Student’s t test, t = 4.05, p = 0.0003). This concentration-dependent increase in tonic current by IL-1b was completely blocked

by IL-1ra (250 ng ml 1 for 30 min, n = 11–22; one-way ANOVA, F(3,61) = 7.92, p = 0.0002; Dunnett’s post hoc test, p < 0.001 compared with control; Figure 3C). Because acute inflammation is associated with increased levels of other key cytokines, including TNF-a and IL-6 (Dantzer et al., 2008; Yirmiya and Goshen, 2011), we examined the effects of these cytokines on the tonic current. No changes in the tonic current were observed when neurons were treated for 20 min with TNF-a (100 ng ml 1 0.9 ± 0.08 pA pF 1, 500 ng ml 1 1.0 ± 0.1 pA pF 1, versus control 1.0 ± 0.06 pA pF 1, n = 10–14; one-way ANOVA, F(2,34) = 0.05, p = 0.95) or IL-6 (10 ng ml 1 0.9 ± 0.05 pA pF 1, versus control 1.0 ± 0.06 pA pF 1, n = 5; Student’s t test, t = 0.50, p = 0.63). Neurons in the hippocampus express a variety of GABAA receptor subtypes (Glykys and Mody, 2007; Luscher et al.,

492 Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors

A

C

Figure 4. Mechanisms for IL-1b-Induced Enhancement of the Tonic Current

B

(A) Treatment with an inhibitor of p38 MAPK, SB203,580 (p38 Anta, 20 mM, 30 min) abolished the effect of IL-1b (20 ng ml 1, 20 min) on the tonic current (left). SB202,474 (SB202, 10 mM, 30 min), an inactive analog of the p38 MAPK inhibitor SB203,580, did not block the enhancing effects of IL-1b on the tonic current (right). (B and C) IL-1b-induced enhancement of the tonic current was not blocked by the JNK antagonist SP600,125 (JNK Anta, 1 mM, 30 min) or the PI3K antagonist LY294,002 (PI3K Anta, 20 mM, 30 min). (B) n = 11–16; one-way ANOVA, F(2,37) = 14.42, p < 0.0001. (C) n = 10–12; one-way ANOVA, F(2,30) = 11.13, p = 0.0002. Dunnett’s post hoc test, ***p < 0.001 compared with control. (D) The surface expression of the a5 subunit in hippocampal slices treated with IL-1b (20 ng ml 1, 40 min) was increased compared with vehicletreated control slices. Data are represented as mean ± SEM. See also Figure S4.

D

2011). To further examine whether the IL-1b-enhanced tonic current was indeed mediated by a5GABAA receptors, we used a combination of pharmacological and genetic approaches. We first studied the effects of L-655,708 (20 nM), which inhibited the tonic current in WT neurons by 56.6% ± 9.2% (n = 7; Figure 3D). This effect size is consistent with the efficacy of L-655,708 for inhibition of a5GABAA receptors (Quirk et al., 1996). Next, we found that IL-1b increased the tonic current revealed by both Bic and L-655,708 recorded in WT neurons, but did not increase the tonic current recorded in Gabra5 / neurons (Figures 3E and 3F). These results suggest that the a5GABAA receptors are necessary for IL-1b enhancement of the tonic current. We next studied whether elevated levels of endogenous IL-1b also enhance the tonic current. Cocultures of microglia and neurons were prepared and treated with the endotoxin LPS (100 ng ml 1 for 12–15 h) to stimulate release of IL-1b, as previously described (Polazzi and Contestabile, 2006). LPS treatment increased the tonic current, an effect that was fully reversed by treating the cocultures with IL-1ra (250 ng ml 1, n = 13–19; one-way ANOVA, F(3,62) = 7.27, p = 0.0003; Newman-Keuls post hoc test, ** p < 0.01, *** p < 0.001 compared with LPS; Figure 3G). However, the tonic current studied in neurons that were cultured alone, in the absence of microglia, was unchanged by LPS (n = 10–14; Student’s t test, t = 0.10, p = 0.92; Figure 3G), which suggests that IL-1b was released from activated microglia in the cocultures (Polazzi and Contestabile, 2006). In the hippocampus, GABAA receptors also generate fast transient inhibitory postsynaptic currents that are predominantly

mediated by GABAA receptors lacking the a5 subunit (Glykys and Mody, 2007). To determine whether IL-1b also enhances synaptic inhibition, we recorded miniature inhibitory postsynaptic currents (mIPSCs) from cultured hippocampal neurons. IL-1b caused a modest reduction in the mIPSCs (Figure S3; Table S1). Collectively, these results suggest that IL-1b preferentially increases the activity of GABAA receptors that generate tonic (but not synaptic) currents. IL-1b Increases the Tonic Current via the p38 MitogenActivated Protein Kinase-Dependent Pathway Next, to identify the predominant signaling pathways by which IL-1b increases the tonic currents, we treated hippocampal neurons with selective kinase inhibitors. SB203,580, an inhibitor of p38 mitogen-activated protein kinase (MAPK), completely blocked the IL-1b-induced increase in the tonic current (n = 10–14; one-way ANOVA, F(3,45) = 4.85, p = 0.005; Dunnett’s post hoc test, p < 0.05 compared with control; Figure 4A), whereas its inactive analog, SB202,474 had no effect (n = 9–11; one-way ANOVA, F(3,35) = 11.10, p < 0.0001; Dunnett’s post hoc test, ** p < 0.01 *** p < 0.001 compared with control; Figure 4A). This result suggested that p38 MAPK mediates the IL-1b-induced increase in tonic current. Inhibitors of c-Jun N-terminal kinases (JNKs; SP600,125) and phosphatidylinositol 3-kinases (PI3Ks; LY294,002) had no effect on IL-1b enhancement of the tonic current, showing that activation of these kinases was not required (Figures 4B and 4C). MAPK signaling pathways are important regulators of GABAA receptor trafficking (Luscher et al., 2011). Thus, we studied whether IL-1b increased the expression of a5GABAA receptors on the surface of neurons. Hippocampal slices were treated with IL-1b (20 ng ml 1 for 40 min), and a quantitative western blot analysis of the biotinylated protein was performed (Abramian

Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors 493

et al., 2010). The surface expression of the a5-subunit protein in IL-1b-treated slices was increased to 157.4% ± 17.6% of the surface protein measured in vehicle-treated control slices (n = 6, one-sample t test, t = 3.27, p = 0.02; Figure 4D). The surface expression of the a1 and a2 subunits was also studied, because these subunits are widely expressed and contribute primarily to synaptic GABAA receptors (Glykys and Mody, 2007; Luscher et al., 2011). IL-1b decreased the surface expression of the a1 subunit but did not modify the surface expression of the a2 subunit (Figure S4). These results are consistent with the observation that IL-1b caused a modest decrease in the amplitude of postsynaptic currents. DISCUSSION The results described here show that inflammation causes memory loss, at least in part, by increasing a tonic inhibitory conductance that is generated by a5GABAA receptors. We observed that genetic or pharmacological inhibition of a5GABAA receptors was sufficient to prevent impairment of contextual fear memory and synaptic plasticity induced by LPS and IL-1b. Further, IL-1b increased the tonic conductance generated by a5GABAA receptors in hippocampal neurons via activation of IL-1 receptors and p38 MAPK-dependent signaling. This increased tonic conductance by IL-1b was most likely due to an increase in the surface expression of a5GABAA receptors. Multiple signaling pathways are activated by IL-1b binding to the IL-1 receptor, including pathways involving p38 MAPK, JNKs, and PI3Ks (O’Neill, 2002). Our results show that IL-1b enhances the tonic inhibitory conductance mediated by a5GABAA receptors primarily via activation of the p38 MAPKdependent pathway. Activation of p38 MAPK by IL-1b has been shown in hippocampal neurons (Srinivasan et al., 2004) and cortical neurons (Li et al., 2003). Also, consistent with our results, others have shown that activation of p38 MAPK mediates the effects of IL-1b on hippocampal LTP (Coogan et al., 1999; Kelly et al., 2003). The increase in tonic current was accompanied by an increase in the surface expression of a5GABAA receptors. Membrane clustering of a5GABAA receptors depends on the presence of radixin, an actin-binding protein that anchors the a5 subunit to cytoskeletal elements (Loebrich et al., 2006). Radixin is targeted by p38 MAPK (Koss et al., 2006), and phosphorylation of radixin could lead to enhanced membrane stability and clustering of a5GABAA receptors at extracellular sites. Alternatively, IL1b-dependent activation of p38 MAPK could lead to increased activity of downstream mediators such as cAMP response element-binding protein (Srinivasan et al., 2004), which may also enhance the surface expression of a5GABAA receptors. The available evidence indicates that IL-1b increases the tonic current primarily by increasing the expression of receptors on the surface rather than causing an increase in the concentration of agonist. Specifically, we studied the tonic current in cultured neurons and slices under ‘‘concentration-clamp’’ conditions, in which the extracellular concentration of GABA was fixed. Under these conditions, IL-1b increased the amplitude of the tonic current. Nevertheless, it is plausible that IL-1b also causes a concurrent increase in ambient levels of GABA in the brain in

addition to increasing the surface expression of a5GABAA receptors. The concentration of GABA in the extracellular space increases in many proinflammatory states, such as stress (de Groote and Linthorst, 2007), stroke (Clarkson et al., 2010), and pain (Kupers et al., 2009). The increase in extracellular GABA could result from enhanced transmitter release and/or decreased reuptake of GABA. It is unlikely that such an increase in GABA induced by IL-1b is due to synaptic vesicular release, because the frequency of mIPSCs was not increased in the current study. Instead, reduced reuptake via GABA transporters (Wu et al., 2007), as well as increased release of GABA from activated glial cells (Lee et al., 2010), may increase GABA levels in the brain. It is of interest that inflammation impairs memory performance in response to context but not to tone, because the former (but not the latter) behavior depends on the hippocampus (Fanselow and Poulos, 2005). This result is consistent with the relatively restricted expression of a5GABAA receptors to the hippocampus (Pirker et al., 2000). In the hippocampus, d-subunit-containing GABAA (dGABAA) receptors are also located in extrasynaptic regions of the neurons, where they generate a tonic inhibitory conductance (Glykys and Mody, 2007). Our results do not rule out the possibility that inflammation modifies other GABAA receptor subtypes, including dGABAA receptors; however, these other subtypes are unlikely to play a role in memory loss. Gabra5 / mice express dGABAA receptors, yet IL-1b failed to increase the tonic current in the hippocampal neurons of these mice. Also, inflammation did not induce memory deficits in Gabra5 / mice. Of note, inflammation impairs synaptic plasticity in the dentate gyrus region, where the expression of a5GABAA receptors is low under baseline conditions (Glykys and Mody, 2007; Kelly et al., 2003). The levels of a5GABAA receptors in the dentate gyrus could be increased during inflammation. In support of this notion, the activity of a5GABAA receptors increases in the frontal cortex after ischemic stroke (Clarkson et al., 2010). The current study has several limitations. The first of these relates to the widely used model that we employed to induce systemic inflammation via intraperitoneal injection of IL-1b and LPS. This experimental model mimics many of the features of acute systemic illness in humans (Yirmiya and Goshen, 2011); however, the method does not restrict the induction of inflammation to the hippocampus. Second, the Gabra5 / mouse does not limit the reduction of expression either regionally or temporally. Also, global deletion of genes for GABAA receptor subunits can lead to compensatory changes in the expression of other subunits or proteins, although this has not been reported for Gabra5 / mice (Collinson et al., 2002; Crestani et al., 2002). Behaviors elicited by pharmacological treatments with L-655,708 and MRK-016 faithfully mimicked behaviors exhibited by Gabra5 / mice. Future studies employing conditional mutagenesis of the Gabra5 gene that is restricted to the CA1 subfield would advance our understanding of the specific role of a5GABAA receptors in the CA1 hippocampus. Finally, although the consequences of injecting various drugs directly into the hippocampus would be of interest, we did not perform such experiments, in part because of the requirement for carefully designed and performed control experiments, given that

494 Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors

inflammation induced by surgery (Cibelli et al., 2010) and treatment with general anesthetics (Zurek et al., 2012) can cause long-term impairment of memory performance. Currently, no treatments are available to reduce the memory deficits associated with inflammation. The results of this study indicate that the attenuation of memory loss by the administration of inverse agonists for a5GABAA receptors is a potential treatment strategy. Because of the restricted expression of a5GABAA receptors in the hippocampus (Pirker et al., 2000), the inverse agonists that inhibit these receptors are generally well tolerated and lack the proconvulsant and anxiogenic properties of nonselective GABAA receptor antagonists (Atack, 2010). Finally, this study may stimulate interest in the effects of inflammation on GABAA receptors that are expressed in tissues outside the nervous system (Watanabe et al., 2002), including the lung and liver. Upregulation of these GABAA receptors may also contribute to organ dysfunction associated with infection, trauma, and immune-related diseases (Dantzer et al., 2008). EXPERIMENTAL PROCEDURES All experimental procedures were approved by the Animal Care Committee of the University of Toronto (Toronto, ON, Canada). In all studies, the experimenter was blinded to the drug treatment and genotype of the mice. For behavioral analysis and studies of hippocampal plasticity, male 3- to 4month-old WT and Gabra5 / mice were used. Experiments were started 3 hr after injection of IL-1b (1 mg kg 1, i.p.) or LPS (125 mg kg 1, i.p.) because cytokine levels, and particularly IL-1b peak at this time point (Hansen et al., 2000; Yirmiya and Goshen, 2011). Pavlovian fear conditioning was used to study memory performance. Each mouse was trained to associate the context and a 3,600 Hz tone lasting 20 s with an electric foot shock (2 s, 0.5 mA). Each sequence was presented three times, separated by 60 s. LTP was induced in the hippocampal slices with a theta burst stimulation, which consisted of 10 stimulus trains at 5 Hz, with each train consisting of four pulses at 100 Hz. Whole-cell voltage-clamp recordings were obtained from hippocampal slices, cell culture, and neuron-microglia coculture. To record the tonic GABAergic current, 5 or 0.5 mM GABA was added to the slices or cultured neurons, respectively. All recordings were performed at a holding potential of 60 mV. A cell-surface biotinylation assay was used to determine the surface expression of GABAA receptor a1, a2, and a5 subunits. Data are represented as the mean ± SEM. A Student’s paired or unpaired t test was used to compare pairs of data. To compare three or more groups, a one-way (drug treatment only) or two-way (drug treatment versus genotype) ANOVA followed by a Dunnett, Newman-Keuls, or Bonferroni post hoc test was used. A one-sample t test was used to compare the surface expression of GABAA receptor subunits. Statistical significance was set at p < 0.05.

MOP-38028 and MOP-79428 to B.A.O. A.A.Z. is supported by a Natural Sciences and Engineering Research Council of Canada Graduate Scholarship. I.L. is supported by a Savoy Foundation Studentship. S.A. is supported by a Canadian Anesthesia Society/Vitaid Residents’ Research Grant from the Canadian Anesthesia Research Foundation. P.A.D. is supported by a grant from the National Institute of Alcoholism and Alcohol Abuse, National Institutes of Health (AA017938). B.A.O. holds a Canada Research Chair. Received: May 27, 2012 Revised: August 8, 2012 Accepted: August 21, 2012 Published online: September 20, 2012 REFERENCES Abramian, A.M., Comenencia-Ortiz, E., Vithlani, M., Tretter, E.V., Sieghart, W., Davies, P.A., and Moss, S.J. (2010). Protein kinase C phosphorylation regulates membrane insertion of GABAA receptor subtypes that mediate tonic inhibition. J. Biol. Chem. 285, 41795–41805. Atack, J.R. (2010). Preclinical and clinical pharmacology of the GABAA receptor a5 subtype-selective inverse agonist a5IA. Pharmacol. Ther. 125, 11–26. Atack, J.R., Maubach, K.A., Wafford, K.A., O’Connor, D., Rodrigues, A.D., Evans, D.C., Tattersall, F.D., Chambers, M.S., MacLeod, A.M., Eng, W.S., et al. (2009). In vitro and in vivo properties of 3-tert-butyl7-(5-methylisoxazol-3-yl)-2-(1-methyl-1H-1,2,4-triazol-5-ylmethoxy)-pyrazolo [1,5-d]-[1,2,4]triazine (MRK-016), a GABAA receptor a5 subtype-selective inverse agonist. J. Pharmacol. Exp. Ther. 331, 470–484. Brickley, S.G., and Mody, I. (2012). Extrasynaptic GABA(A) receptors: their function in the CNS and implications for disease. Neuron 73, 23–34. Burgard, E.C., Tietz, E.I., Neelands, T.R., and Macdonald, R.L. (1996). Properties of recombinant g-aminobutyric acid A receptor isoforms containing the a 5 subunit subtype. Mol. Pharmacol. 50, 119–127. Caraiscos, V.B., Elliott, E.M., You-Ten, K.E., Cheng, V.Y., Belelli, D., Newell, J.G., Jackson, M.F., Lambert, J.J., Rosahl, T.W., Wafford, K.A., et al. (2004). Tonic inhibition in mouse hippocampal CA1 pyramidal neurons is mediated by a5 subunit-containing g-aminobutyric acid type A receptors. Proc. Natl. Acad. Sci. USA 101, 3662–3667. Cheng, V.Y., Martin, L.J., Elliott, E.M., Kim, J.H., Mount, H.T., Taverna, F.A., Roder, J.C., Macdonald, J.F., Bhambri, A., Collinson, N., et al. (2006). a5GABAA receptors mediate the amnestic but not sedative-hypnotic effects of the general anesthetic etomidate. J. Neurosci. 26, 3713–3720. Cibelli, M., Fidalgo, A.R., Terrando, N., Ma, D., Monaco, C., Feldmann, M., Takata, M., Lever, I.J., Nanchahal, J., Fanselow, M.S., and Maze, M. (2010). Role of interleukin-1b in postoperative cognitive dysfunction. Ann. Neurol. 68, 360–368. Clarkson, A.N., Huang, B.S., Macisaac, S.E., Mody, I., and Carmichael, S.T. (2010). Reducing excessive GABA-mediated tonic inhibition promotes functional recovery after stroke. Nature 468, 305–309.

SUPPLEMENTAL INFORMATION Supplemental Information includes Extended Experimental Procedures, four figures, and one table and can be found with this article online at http://dx. doi.org/10.1016/j.celrep.2012.08.022. LICENSING INFORMATION This is an open-access article distributed under the terms of the Creative Commons Attribution-Noncommercial-No Derivative Works 3.0 Unported License (CC-BY-NC-ND; http://creativecommons.org/licenses/by-nc-nd/3.0/ legalcode).

Collinson, N., Kuenzi, F.M., Jarolimek, W., Maubach, K.A., Cothliff, R., Sur, C., Smith, A., Otu, F.M., Howell, O., Atack, J.R., et al. (2002). Enhanced learning and memory and altered GABAergic synaptic transmission in mice lacking the a 5 subunit of the GABAA receptor. J. Neurosci. 22, 5572–5580. Coogan, A.N., O’Neill, L.A., and O’Connor, J.J. (1999). The P38 mitogen-activated protein kinase inhibitor SB203580 antagonizes the inhibitory effects of interleukin-1b on long-term potentiation in the rat dentate gyrus in vitro. Neuroscience 93, 57–69.

ACKNOWLEDGMENTS

Crestani, F., Keist, R., Fritschy, J.M., Benke, D., Vogt, K., Prut, L., Blu¨thmann, H., Mo¨hler, H., and Rudolph, U. (2002). Trace fear conditioning involves hippocampal a5 GABA(A) receptors. Proc. Natl. Acad. Sci. USA 99, 8980–8985.

We thank Ella Czerwinska for her assistance with the cell cultures. This work was supported by Canadian Institutes of Health Research operating grants

Dantzer, R., O’Connor, J.C., Freund, G.G., Johnson, R.W., and Kelley, K.W. (2008). From inflammation to sickness and depression: when the immune system subjugates the brain. Nat. Rev. Neurosci. 9, 46–56.

Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors 495

de Groote, L., and Linthorst, A.C. (2007). Exposure to novelty and forced swimming evoke stressor-dependent changes in extracellular GABA in the rat hippocampus. Neuroscience 148, 794–805.

Martin, L.J., Oh, G.H., and Orser, B.A. (2009). Etomidate targets a5 g-aminobutyric acid subtype A receptors to regulate synaptic plasticity and memory blockade. Anesthesiology 111, 1025–1035.

Di Filippo, M., Sarchielli, P., Picconi, B., and Calabresi, P. (2008). Neuroinflammation and synaptic plasticity: theoretical basis for a novel, immune-centred, therapeutic approach to neurological disorders. Trends Pharmacol. Sci. 29, 402–412.

Martin, L.J., Zurek, A.A., MacDonald, J.F., Roder, J.C., Jackson, M.F., and Orser, B.A. (2010). a5GABAA receptor activity sets the threshold for longterm potentiation and constrains hippocampus-dependent memory. J. Neurosci. 30, 5269–5282.

Fanselow, M.S., and Poulos, A.M. (2005). The neuroscience of mammalian associative learning. Annu. Rev. Psychol. 56, 207–234.

O’Neill, L.A. (2002). Signal transduction pathways activated by the IL-1 receptor/toll-like receptor superfamily. Curr. Top. Microbiol. Immunol. 270, 47–61.

Fleischmann, R.M., Tesser, J., Schiff, M.H., Schechtman, J., Burmester, G.R., Bennett, R., Modafferi, D., Zhou, L., Bell, D., and Appleton, B. (2006). Safety of extended treatment with anakinra in patients with rheumatoid arthritis. Ann. Rheum. Dis. 65, 1006–1012. Glykys, J., and Mody, I. (2007). Activation of GABAA receptors: views from outside the synaptic cleft. Neuron 56, 763–770. Hansen, M.K., Nguyen, K.T., Fleshner, M., Goehler, L.E., Gaykema, R.P., Maier, S.F., and Watkins, L.R. (2000). Effects of vagotomy on serum endotoxin, cytokines, and corticosterone after intraperitoneal lipopolysaccharide. Am. J. Physiol. Regul. Integr. Comp. Physiol. 278, R331–R336. Ju, Y.H., Guzzo, A., Chiu, M.W., Taylor, P., Moran, M.F., Gurd, J.W., MacDonald, J.F., and Orser, B.A. (2009). Distinct properties of murine a 5 g-aminobutyric acid type a receptors revealed by biochemical fractionation and mass spectroscopy. J. Neurosci. Res. 87, 1737–1747. Kelly, A., Vereker, E., Nolan, Y., Brady, M., Barry, C., Loscher, C.E., Mills, K.H., and Lynch, M.A. (2003). Activation of p38 plays a pivotal role in the inhibitory effect of lipopolysaccharide and interleukin-1 b on long term potentiation in rat dentate gyrus. J. Biol. Chem. 278, 19453–19462. Kipnis, J., Derecki, N.C., Yang, C., and Scrable, H. (2008). Immunity and cognition: what do age-related dementia, HIV-dementia and ‘chemo-brain’ have in common? Trends Immunol. 29, 455–463. Koss, M., Pfeiffer, G.R., 2nd, Wang, Y., Thomas, S.T., Yerukhimovich, M., Gaarde, W.A., Doerschuk, C.M., and Wang, Q. (2006). Ezrin/radixin/moesin proteins are phosphorylated by TNF-a and modulate permeability increases in human pulmonary microvascular endothelial cells. J. Immunol. 176, 1218– 1227. Kupers, R., Danielsen, E.R., Kehlet, H., Christensen, R., and Thomsen, C. (2009). Painful tonic heat stimulation induces GABA accumulation in the prefrontal cortex in man. Pain 142, 89–93. Lee, S., Yoon, B.E., Berglund, K., Oh, S.J., Park, H., Shin, H.S., Augustine, G.J., and Lee, C.J. (2010). Channel-mediated tonic GABA release from glia. Science 330, 790–796. Li, Y., Liu, L., Barger, S.W., and Griffin, W.S. (2003). Interleukin-1 mediates pathological effects of microglia on tau phosphorylation and on synaptophysin synthesis in cortical neurons through a p38-MAPK pathway. J. Neurosci. 23, 1605–1611. Loebrich, S., Ba¨hring, R., Katsuno, T., Tsukita, S., and Kneussel, M. (2006). Activated radixin is essential for GABAA receptor a5 subunit anchoring at the actin cytoskeleton. EMBO J. 25, 987–999. Luscher, B., Fuchs, T., and Kilpatrick, C.L. (2011). GABAA receptor traffickingmediated plasticity of inhibitory synapses. Neuron 70, 385–409. Lynch, A.M., Walsh, C., Delaney, A., Nolan, Y., Campbell, V.A., and Lynch, M.A. (2004). Lipopolysaccharide-induced increase in signalling in hippocampus is abrogated by IL-10—a role for IL-1 beta? J. Neurochem. 88, 635–646. Lynch, M.A. (2004). Long-term potentiation and memory. Physiol. Rev. 84, 87–136.

Pirker, S., Schwarzer, C., Wieselthaler, A., Sieghart, W., and Sperk, G. (2000). GABA(A) receptors: immunocytochemical distribution of 13 subunits in the adult rat brain. Neuroscience 101, 815–850. Polazzi, E., and Contestabile, A. (2006). Overactivation of LPS-stimulated microglial cells by co-cultured neurons or neuron-conditioned medium. J. Neuroimmunol. 172, 104–111. Quirk, K., Blurton, P., Fletcher, S., Leeson, P., Tang, F., Mellilo, D., Ragan, C.I., and McKernan, R.M. (1996). [3H]L-655,708, a novel ligand selective for the benzodiazepine site of GABAA receptors which contain the a 5 subunit. Neuropharmacology 35, 1331–1335. Serantes, R., Arnalich, F., Figueroa, M., Salinas, M., Andre´s-Mateos, E., Codoceo, R., Renart, J., Matute, C., Cavada, C., Cuadrado, A., and Montiel, C. (2006). Interleukin-1b enhances GABAA receptor cell-surface expression by a phosphatidylinositol 3-kinase/Akt pathway: relevance to sepsis-associated encephalopathy. J. Biol. Chem. 281, 14632–14643. Srinivasan, D., Yen, J.H., Joseph, D.J., and Friedman, W. (2004). Cell typespecific interleukin-1b signaling in the CNS. J. Neurosci. 24, 6482–6488. Stellwagen, D., Beattie, E.C., Seo, J.Y., and Malenka, R.C. (2005). Differential regulation of AMPA receptor and GABA receptor trafficking by tumor necrosis factor-alpha. J. Neurosci. 25, 3219–3228. Terrando, N., Eriksson, L.I., Ryu, J.K., Yang, T., Monaco, C., Feldmann, M., Jonsson Fagerlund, M., Charo, I.F., Akassoglou, K., and Maze, M. (2011). Resolving postoperative neuroinflammation and cognitive decline. Ann. Neurol. 70, 986–995. Trompet, S., de Craen, A.J., Slagboom, P., Shepherd, J., Blauw, G.J., Murphy, M.B., Bollen, E.L., Buckley, B.M., Ford, I., Gaw, A., et al; PROSPER Group. (2008). Genetic variation in the interleukin-1 b-converting enzyme associates with cognitive function. The PROSPER study. Brain 131, 1069–1077. Viviani, B., Bartesaghi, S., Gardoni, F., Vezzani, A., Behrens, M.M., Bartfai, T., Binaglia, M., Corsini, E., Di Luca, M., Galli, C.L., and Marinovich, M. (2003). Interleukin-1b enhances NMDA receptor-mediated intracellular calcium increase through activation of the Src family of kinases. J. Neurosci. 23, 8692–8700. Watanabe, M., Maemura, K., Kanbara, K., Tamayama, T., and Hayasaki, H. (2002). GABA and GABA receptors in the central nervous system and other organs. Int. Rev. Cytol. 213, 1–47. Wu, Y., Wang, W., Dı´ez-Sampedro, A., and Richerson, G.B. (2007). Nonvesicular inhibitory neurotransmission via reversal of the GABA transporter GAT-1. Neuron 56, 851–865. Yirmiya, R., and Goshen, I. (2011). Immune modulation of learning, memory, neural plasticity and neurogenesis. Brain Behav. Immun. 25, 181–213. Zurek, A.A., Bridgwater, E.M., and Orser, B.A. (2012). Inhibition of a5 g-Aminobutyric acid type A receptors restores recognition memory after general anesthesia. Anesth. Analg. 114, 845–855.

496 Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors

Supplemental Information EXTENDED EXPERIMENTAL PROCEDURES Experimental Animals All experimental procedures and protocols were approved by the Animal Care Committee of the University of Toronto (Toronto, Ontario, Canada). Mice were obtained from two sources. Swiss White mice were used to prepare hippocampal cell cultures (Charles River, Wilmington, MA, USA). Male a5GABAA receptor null mutant (Gabra5/) and wild-type (WT) mice (C57BL/6/SvEv129) were used for behavioral studies, slice recordings and hippocampal cell cultures. They were obtained from the authors’ breeding colony at the University of Toronto. The generation, genotyping, and characterization of Gabra5/ mice have been previously described (Collinson et al., 2002). The experimenter was blinded to the drug treatment and genotype of the mice for all studies. Fear Conditioning Pavlovian fear-conditioning was used for behavioral analysis (Martin et al., 2010; Wehner and Radcliffe, 2004). Male, 3- to 4-monthold WT and Gabra5/ mice were handled in 5 min epochs every day, for 5 days before behavioral training. Mice were trained 3 hr after injection of IL-1b (1 mg kg1, i.p., with 0.1% bovine serum albumin in phosphate buffered saline as the vehicle control) or LPS (125 mg kg1, i.p., with physiological saline as the vehicle control). During training, each mouse was allowed to explore the chamber for 180 s. A 3600 Hz tone, created by a frequency generator, amplified to 70 dB and lasting 20 s was then presented. The last 2 s of each auditory tone were paired with an electric footshock (2 s, 0.5 mA). Each of these sequences was presented three times, separated by 60 s. For some mice, the inverse agonists L-655,708 (0.35 or 0.5 mg kg1, i.p.) or MRK-016 (3 mg kg1, i.p.) or vehicle control (10% DMSO, i.p.) was injected 30 min before training. On day 2, 24 hr after the training session, contextual fear was assessed by placing the mouse in the original context for 8 min and measuring the percentage of total time that the mouse spent freezing. On day 3, the conditioning chamber was modified to measure the freezing response to the conditioned auditory tone (cued fear conditioning). Mice were monitored for 180 s for freezing to the modified context to rule out contextual influences. After the monitoring period, the auditory tone was presented continuously for 5 min and movement of the mouse was monitored. The percentage of time that each mouse spent freezing was determined using FreezeView software (Version 2.26, Actimetrics Inc, Wilmette, IL, USA). Synaptic Plasticity in Hippocampal Slices The long-term potentiation (LTP) of evoked field postsynaptic potentials (fPSPs) was studied using previously described methods (Martin et al., 2010). Male 3- to 4-month-old WT or Gabra5/ mice were injected with either LPS (125 mg kg1, i.p.) or vehicle control (physiological saline, i.p.), and hippocampal slices (350 mm) were prepared 3 hr later. fPSPs were recorded at room temperature (21 – 23 C). Baseline stimulation frequency was 0.05 Hz, and the stimulus intensity was adjusted to evoke a half-maximal fPSP amplitude. LTP was induced in the slices with theta burst stimulation, which consisted of 10 stimulus trains at 5 Hz, with each train including 4 pulses at 100 Hz. The fPSPs were monitored for 10 min before TBS and 60 min after TBS. Whole-Cell Voltage-Clamp Recordings in Hippocampal Slices For whole-cell voltage clamp recordings, hippocampal slices were prepared from 3 – 4 week-old WT mice with a mixed genetic background (C57BL/6/SvEv129). Mice were injected with IL-1b (1 mg kg1, i.p.) or vehicle (0.1% bovine serum albumin in phosphate buffered saline, i.p.) 2–3 hr before sacrifice. Mice were decapitated under isoflurane anesthesia and their brains were quickly removed and placed in ice-cold, oxygenated (95% O2, 5% CO2) artificial cerebrospinal fluid (ACSF, composition in mM: 124 NaCl, 3 KCl, 1.3 MgCl2, 2.6 CaCl2, 1.25 NaH2PO4, 26 NaHCO3, and 10 D-glucose) with the osmolarity adjusted to 300 – 310 mOsm. Coronal brain slices (350 mm) were prepared with a VT1200S vibratome (Leica, Deerfield, IL, USA). After a recovery period of 1 hr in oxygenated ACSF, the slices were transferred to a submersion-type recording chamber. To record tonic GABAergic currents 5 mM GABA was added to ACSF. The amplitude of the tonic current was calculated as the difference between the holding current measured before and after the application of the GABAA receptor competitive antagonist, bicuculline (10 mM) or L-655,708 (20 nM). The holding current was calculated from segments containing no synaptic currents (Brickley et al., 1996). Patch-pipettes (3–5 MU) were fabricated from borosilicate glass capillary tubing and filled with a solution containing (in mM): 140 CsCl, 10 HEPES, 11 EGTA, 4 Mg2ATP, 2 MgCl2, 1 CaCl2, 2 TEA (pH 7.3 with CsOH, 290 – 300 mOsm). All experiments were conducted at room temperature and whole-cell currents were recorded using the Multiclamp 700B amplifier (Molecular Devices, Sunnyvale, CA, USA) controlled with pClamp 9.0 software via a Digidata 1322 interface (Molecular Devices, Sunnyvale, CA, USA). Currents were sampled at 10 kHz and filtered at 2 kHz by using an eight-pole low-pass Bessel filter. All cells were recorded at a holding potential of 60 mV. Cell Culture Cultures of embryonic hippocampal neurons were prepared as previously described (Bonin et al., 2007). Briefly, fetal pups (embryonic day 18) were removed from mice euthanized by cervical dislocation. The hippocampi were dissected from each fetus and placed in an ice-cooled culture dish. Neurons were then dissociated by mechanical titration using two Pasteur pipettes (tip diameter, 150– 200 mm) and plated on 35-mm culture dishes at a density of approximately 1 3 106 cells ml1. The culture dishes were coated with collagen or poly-D-lysine (Sigma-Aldrich Co., St. Louis, MO, USA). For the first 5 days in vitro (DIV), cells were maintained in minimal essential media (MEM) supplemented with 10% fetal bovine serum (FBS) and 10% horse serum (Life Technologies, Grand Island, NY, USA). The neurons were cultured at 37 C in a 5% CO2-95% air environment. After the background cells had grown to confluence, Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors S1

0.1 ml of a mixture of 4 mg 5-fluorodeoxyuridine and 10 mg uridine in 20 ml MEM was added to the extracellular solution to reduce the number of dividing cells. Subsequently, the media was supplemented with 10% horse serum and changed every 3 or 4 days. Cells were maintained in culture for 14 to 20 days before use. For all reported results, data were acquired from cells from at least three different dissections. Neuron-Microglia Coculture Primary cultures of hippocampal neurons were prepared from fetal pups (embryonic day 18). In brief, cells were dissociated from hippocampi and plated on 35-mm dishes in neurobasal medium supplemented with 0.5% FBS, 2% B-27 supplement, 0.5 mM Lglutamine, 25 mM glutamic acid and 0.5 mM Na+ pyruvate. After 24 – 48 hr, FUDR (5 mM Uridine and 5 mM 5-Fluoro-20 -deoxy-uridine) was added to avoid glial proliferation. The cultures were maintained at 37 C in a humidified atmosphere of 5% CO2 and the medium was changed every 3 or 4 days. Cultures of hippocampal neurons were maintained in culture for 12-16 days before use. Mixed microglial cultures were prepared from embryonic cerebral cortex of 15- or 16-day-old mice. Cells were cultured in glial culture media (MEM supplemented with 10% fetal bovine serum) on 60-mm dishes at 37 C in a 5% CO2 incubator; the medium was changed every 3 or 4 days. Once confluence was achieved, microglia were separated from the mixed glial culture (typically at 10–14 days) by gentle shaking (200 rpm for 2 hr at 37 C) and were then collected by centrifugation. The pellet was suspended in neurobasal medium and applied directly over cultured hippocampal neurons at 10 – 14 days in vitro. To induce microglial activation, primary microglia-enriched cultures were treated with LPS (100 ng ml1) overnight (12–15 h) before electrophysiological recordings. Whole-Cell Voltage-Clamp Recordings in Cultured Neurons Prior to electrophysiological recordings, cultured neurons were rinsed with extracellular recording solution containing (in mM): 140 NaCl, 2 CaCl2, 1 MgCl2, 5.4 KCl, 25 N-2-hydroxy-ethylpiperazine-N’-2-ethanesulphonic acid (HEPES), 28 glucose (pH 7.4, 325–335 mOsm). 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX, 10 mM) and (2R)- amino-5-phosphonovaleric acid (APV, 40 mM) were added to the extracellular solution to block ionotropic glutamate receptors and tetrodotoxin (0.3 mM) was used to block voltage-dependent sodium channels. To record tonic GABAergic currents, 0.5 mM GABA was added to the extracellular solution and the GABAA receptor competitive antagonist, bicuculline (100 mM) or L-655,708 (20 nM) was applied to reveal the tonic current. Patch-pipettes (3–4 MU) were fabricated from borosilicate glass capillary tubing and filled with an intracellular solution containing (in mM): 140 CsCl, 10 HEPES, 11 EGTA, 4 K2ATP, 2 MgCl2, 1 CaCl2 and 2 TEA (pH 7.3 with CsOH, 285–295 mOsm). All experiments were conducted at room temperature and whole-cell currents were recorded using the Axopatch 200A amplifier (Molecular Devices, Sunnyvale, CA) controlled with pClamp 8.0 software via a Digidata 1200 interface (Molecular Devices, Sunnyvale, CA). Membrane capacitance was measured using the membrane test protocol in pClamp 8.0. Currents were sampled at 10 kHz and filtered at 2 kHz by using an eight-pole low-pass Bessel filter. All cells were recorded at a holding potential of 60 mV. The extracellular solution was applied to neurons by a computer-controlled multi-barreled perfusion system (SF-77B; Warner Instruments, Hamden, CT, USA). Hippocampal Slice Cell Surface Biotinylation Assay Hippocampal slices (350 mm thick) from 10 – 14 week-old C57BL/6 mice were prepared with a vibrotome (Leica VT1000S, Deerfield, IL, USA) and placed in ice cold oxygenated ACSF containing (in mM): 124 NaCl, 3 KCl, 25 NaHCO3, 2 MgSO4, 2 CaCl2, 1.1 NaH2PO4, and 10 glucose, pH 7.4. The slices were transferred individually to oxygenated ACSF (95% O2, 5% CO2), and equilibrated in a 30 C water bath for 1 hr. Slices were treated with IL-1b (20 ng ml1, 40 min). After drug treatment, slices were placed on ice and incubated for 30 min with 1 mg ml1 NHS-SS-biotin (Thermo Scientific, Rockford, IL, USA) dissolved in ACSF. Excess biotin was removed by washing slices three times with ice cold PBS and lysis buffer, which contained the following (in mM): 20 Tris-HCl (pH 8.0), 150 NaCl, 5 EDTA, 10 NaF, 2 Na3VO4, 10 sodium pyrophosphate, 1% Triton X-100, and 0.1% sodium dodecyl sulfate (SDS). In addition, the following protease inhibitors were added (in mg ml1): 250 4-(2-aminoethyl) benzenesulfonyl fluoride hydrochloride (AEBSF), 10 leupeptin, 1 pepstatin, and 10 antipain. Insoluble material was removed by centrifugation. The supernatant lysates were incubated with NeutrAvidin beads (Thermo Scientific, Rockford, IL, USA) for 18–24 hr at 4 C. Bound material was eluted with sample buffer and subjected to SDS-PAGE analysis. A western blot analysis with anti-GABAA receptor a1 antibodies (PhosphoSolutions, Aurora, CO, USA or Millipore, Billerica, MA, USA), a2 and a5 antibodies (PhosphoSolutions, Aurora, CO, USA or Abcam, Cambridge, MA, USA) was performed. Blots were quantified using the CCD-based FujiFilm LAS300 system (FujiFilm Corporation, Tokyo, Japan) or the Kodak Image Station 2000R (Kodak, Glendale, CA, USA). Drugs and Other Chemicals Mouse recombinant IL-1b and IL-1ra were obtained from R&D Systems (Minneapolis, MN, USA). Lipopolysaccharide was obtained from Sigma-Aldrich Co. (St. Louis, MO, USA). Other chemicals including EGTA, HEPES and KCl were purchased from Fluka Biochemika (Bucks, Switzerland). DMSO was purchased from Fischer Scientific (Fairlawn, NJ, USA), TTX from Precision Biochemicals, Inc. (Vancouver, BC, Canada) and NaCl, CaCl2, glucose and NaOH from BDH Inc. (Toronto, ON, Canada). All other compounds were purchased from Tocris Bioscience (Bristol, UK). S2 Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors

Statistical Analyses Data are represented as mean ± SEM. Student’s paired or unpaired t test was used to compare pairs of data. For comparing three or more groups, one-way (drug treatment only) or two-way (drug treatment versus genotype) analysis of variance (ANOVA) followed by the Dunnett, Newman–Keuls or Bonferroni post hoc test was used. One sample t test was used to compare surface expression of GABAA receptor subunits. Cumulative distributions of the amplitude and frequency of mIPSCs were compared using the Kolmogorov-Smirnov test. Statistical significance was set at p < 0.05. SUPPLEMENTAL REFERENCES Bonin, R.P., Martin, L.J., MacDonald, J.F., and Orser, B.A. (2007). a5GABAA receptors regulate the intrinsic excitability of mouse hippocampal pyramidal neurons. J. Neurophysiol. 98, 2244–2254. Brickley, S.G., Cull-Candy, S.G., and Farrant, M. (1996). Development of a tonic form of synaptic inhibition in rat cerebellar granule cells resulting from persistent activation of GABAA receptors. J. Physiol. 497, 753–759. Wehner, J.M., and Radcliffe, R.A. (2004). Cued and contextual fear conditioning in mice. Curr. Protoc. Neurosci. Chapter 8, Unit 8.5C.

Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors S3

Figure S1. LPS Did Not Affect the Freezing Scores for Cued Fear Memory in Both WT and Gabra5/ Mice, Related to Figure 1 Data are represented as mean ± SEM; two-way ANOVA; effect of LPS: F(1,53) = 0.75, p = 0.39; effect of genotype: F(1,53) = 0.16, p = 0.69; effect of interaction: F(1,53) = 0.61, p = 0.44; n = 13–15, N.S.: nonsignificant result.

S4 Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors

Figure S2. LPS-Induced Impairment of LTP in Slices from WT Mice Was Mediated by IL-1b, Related to Figure 2 (A) Incubation of slices with IL-1ra (100 ng ml1) for 1 hr abolished the LPS-induced impairment of LTP in slices from WT mice. (B) Quantified data for results shown in (A). Data are represented as mean ± SEM; n = 9–10, one-way ANOVA, F(2,26) = 7.82, p = 0.002; Newman-Keuls post hoc test, *p < 0.05, **p < 0.01, compared with LPS.

Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors S5

Figure S3. The Amplitude of Inhibitory Synaptic Currents Was Slightly Decreased by IL-1b in Cultured Hippocampal Neurons, Related to Figure 3 (A) GABAA receptor-mediated mIPSCs were modestly inhibited by treatment with IL-1b (20 ng ml1), and this effect could be blocked by IL-1ra (250 ng ml1). (B) Traces were averaged from 314–474 individual mIPSCs. (C) Cumulative amplitude and frequency distributions of mIPSCs show that the amplitude was slightly inhibited by IL-1b, an effect abolished by IL-1ra (left). p < 0.001 for control versus IL-1b, p = 0.16 for control versus IL-1ra + IL-1b. The frequency of mIPSCs was not affected by IL-1b (right). p = 0.22 for control versus IL-1b. Kolmogorov-Smirnov test.

S6 Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors

Figure S4. Surface Expression of GABAA Receptor Subunits, Related to Figure 4

(A and B) Expression of the a1 subunit was reduced, whereas that of the a2 subunit was unchanged after IL-1b treatment (20 ng ml1, 40 min). Data are represented as mean ± SEM. A one-sample t test was used. (A) n = 10, t = 3.05, p = 0.014. (B) n = 3, t = 1.04, p = 0.41.

Cell Reports 2, 488–496, September 27, 2012 ª2012 The Authors S7