causal evidence for the role of rem ... - its.caltech.edubi250c/papers/boyce-2016.pdfnov 12,...

6
21. M. K. Pető, S. Mukhopadhyay, K. A. Kelley, Earth Planet. Sci. Lett. 369370, 1323 (2013). 22. M. G. Jackson, S. B. Shirey, E. H. Hauri, M. D. Kurz, H. Rizo, Geochim. Cosmochim. Acta 10.1016/j.gca.2016.02.011 (2016). 23. J. P. Brandenburg, E. H. Hauri, P. E. van Keken, C. J. Ballentine, Earth Planet. Sci. Lett. 276,113 (2008). 24. E. J. Garnero, A. K. McNamara, Science 320, 626628 (2008). 25. M. Horan, R. J. Walker, J. W. Morgan, J. N. Grossman, A. E. Rubin, Chem. Geol. 196, 2742 (2003). 26. T. J. Ireland, R. J. Walker, M. O. Garcia, Chem. Geol. 260, 112128 (2009). 27. W. B. Tonks, H. J. Melosh, J. Geophys. Res. Planets 98, 53195333 (1993). ACKNOWLEDGMENTS We thank T. Mock for assistance with mass spectrometry. This work was improved by helpful discussions with J. ONeil, S. Shirey, and B. Wood. We thank S. Shirey, who provided sample VE-32. We also appreciate the helpful comments and suggestions from three anonymous reviewers. We thank M. Garçon, who developed the four-step 142 Nd acquisition method. This work was supported by NSF grant EAR-1265169 (Cooperative Studies of the Earths Deep Interior program) to R.J.W. and grant EAR-1250419 to S.M. All data are available in the main manuscript and supplementary materials. SUPPLEMENTARY MATERIALS www.sciencemag.org/content/352/6287/809/suppl/DC1 Materials and Methods Figs. S1 to S6 Tables S1 to S5 References (2856) 12 November 2015; accepted 5 April 2016 10.1126/science.aad8563 SLEEP RESEARCH Causal evidence for the role of REM sleep theta rhythm in contextual memory consolidation Richard Boyce, 1 Stephen D. Glasgow, 2 Sylvain Williams, 2 *Antoine Adamantidis 2,3 *Rapid eye movement sleep (REMS) has been linked with spatial and emotional memory consolidation. However, establishing direct causality between neural activity during REMS and memory consolidation has proven difficult because of the transient nature of REMS and significant caveats associated with REMS deprivation techniques. In mice, we optogenetically silenced medial septum g-aminobutyric acidreleasing (MS GABA ) neurons, allowing for temporally precise attenuation of the memory-associated theta rhythm during REMS without disturbing sleeping behavior. REMS-specific optogenetic silencing of MS GABA neurons selectively during a REMS critical window after learning erased subsequent novel object place recognition and impaired fear-conditioned contextual memory. Silencing MS GABA neurons for similar durations outside REMS episodes had no effect on memory. These results demonstrate that MS GABA neuronal activity specifically during REMS is required for normal memory consolidation. T he physiological function of rapid eye move- ment sleep (REMS) is unclear (1). Evidence linking REMS to aspects of memory con- solidation in mammals has been obtained using techniques such as statistical corre- lation, pharmacology, and REMS deprivation (2, 3). However, whether REMS has a direct role in learning and memory remains contro- versial; correlative studies are not definitive, REMS has a transient pattern of occurrence that prevents REMS-selective pharmacological manipulation, and REMS deprivation has crit- ical caveats that are difficult to fully control for (4, 5). During REMS in mice and rats, a promi- nent ~7-Hz theta oscillation is observed in lo- cal field potential (LFP) recordings from cortical structures, including the hippocampus (6, 7). Hippocampal theta rhythms during REMS may contribute to memory consolidation by pro- viding a mechanism for strengthening place cells formed during prior wakefulness (8, 9). Theta rhythm generation requires an intact medial septum (MS) (10, 11), although the MS is not involved in REMS generation itself (12, 13). MS g-aminobutyric acidreleasing (MS GABA ) neurons project to the hippocampus, prob- ably pacing the hippocampal theta rhythm during REMS (1416). In mice, we therefore used optogenetics to silence MS GABA neurons and reduce theta activity selectively during REMS, without disturbing sleeping behavior, to determine whether intact MS GABA neural activity during REMS is important for memory consolidation. Adeno-associated virus (AAVdj) encoding Arc- haerhodopsin fused to an enhanced yellow fluo- rescent protein (eYFP) reporter (ArchT-eYFP) was injected into the MS of VGAT::Cre mice (Fig. 1A, top). The resulting ArchT-eYFP expression was ~95% specific for MS GABA neurons (Fig. 1A, bottom, and fig. S1A), stable, and localized to the MS and the diagonal band of Broca (DBB) region for at least 3 months after injection. MS GABA neural projections were observed throughout the hippo- campus (Fig. 1B). Whole-cell voltage and current clamp record- ings of ArchT-eYFPexpressing MS neurons in acute brain slices (fig. S1B) revealed hyperpolar- ization (-39.9 ± 6.6 mV) and outward current (293.9 ± 69.2 pA) upon 594-nm light exposure (fig. S1B). Single-unit recordings in behaving transfected mice (fig. S1C) confirmed that photo- inhibition during REMS, non-REM sleep (NREMS), and wakefulness rapidly produced a potent and reversible reduction in spiking of putative MS GABA neurons (fig. S1D). We next tested the effect of silencing MS GABA neurons during REMS in freely behaving mice. Photoinhibition with constant light pulses deliv- ered to the MS in ArchT-eYFPexpressing mice (ArchT mice) resulted in significantly (65.3 ± 5.6%) reduced theta power measured from dorsal hippo- campal area CA1 LFP (CA1LFP) recording (Fig. 1D, top). No other frequency bands were affected, and the spectral profile of the CA1LFP returned to baseline levels almost immediately upon release of MS GABA neurons from photoinhibition (Fig. 1D, top, and Fig. 2A). Current source density (CSD) analysis revealed that reduced theta power was present in all layers of dorsal hippocampal CA1 (Fig. 2B). Light pulses delivered to the MS of mice only expressing eYFP in MS GABA neurons (YFP control mice) did not affect CA1LFP power (Fig. 1D, bottom), ruling out light as a potential confounding factor in these results. Inhibition of MS GABA neurons did not perturb sleeping behav- ior (Fig. 1D, top), and the probability of state transition during REMS in ArchT mice was un- altered relative to YFP control mice (n = 30 ArchT mice, n = 19 YFP control mice; P = 0.63, unpaired Students t test). We optogenetically silenced MS GABA neurons selectively during each REMS episode after ac- quisition of a novel object place recognition (NOPR) task (Fig. 3A). Mice showed no preference for either object during initial object habituation [day 1 (D1), task acquisition] (Fig. 3A, right). After acquisition, EEG/CA1LFP/EMG (EEG, elec- troencephalogram; EMG, electromyogram) ac- tivity was monitored for 4 hours. Upon entry into REMS, mice in the ArchT or YFP control group had light continuously delivered to the MS until transition to another state occurred, at which time light delivery ceased until subsequent REMS was detected (Fig. 3B). A group of ArchT-eYFP expressing mice that never received light (ArchT control) served as a baseline control for ArchT- eYFP transfection. To determine whether REMS was a critical factor in our results, a final group 812 13 MAY 2016 VOL 352 ISSUE 6287 sciencemag.org SCIENCE 1 Integrated Program in Neuroscience, McGill University, Montreal, Quebec, Canada. 2 Department of Psychiatry, McGill University, Montreal, Quebec, Canada. 3 Department of Neurology and Department of Clinical Research, Inselspital University Hospital, University of Bern, Bern, Switzerland. *Corresponding author. Email: [email protected] (S.W.); [email protected] (A.A.) These authors contributed equally to this work. RESEARCH | REPORTS on April 23, 2018 http://science.sciencemag.org/ Downloaded from on April 23, 2018 http://science.sciencemag.org/ Downloaded from on April 23, 2018 http://science.sciencemag.org/ Downloaded from on April 23, 2018 http://science.sciencemag.org/ Downloaded from

Upload: vuhanh

Post on 01-May-2018

220 views

Category:

Documents


2 download

TRANSCRIPT

21. M. K. Pető, S. Mukhopadhyay, K. A. Kelley, Earth Planet. Sci.Lett. 369–370, 13–23 (2013).

22. M. G. Jackson, S. B. Shirey, E. H. Hauri, M. D. Kurz, H. Rizo,Geochim. Cosmochim. Acta 10.1016/j.gca.2016.02.011 (2016).

23. J. P. Brandenburg, E. H. Hauri, P. E. van Keken, C. J. Ballentine,Earth Planet. Sci. Lett. 276, 1–13 (2008).

24. E. J. Garnero, A. K. McNamara, Science 320, 626–628(2008).

25. M. Horan, R. J. Walker, J. W. Morgan, J. N. Grossman,A. E. Rubin, Chem. Geol. 196, 27–42 (2003).

26. T. J. Ireland, R. J. Walker, M. O. Garcia, Chem. Geol. 260,112–128 (2009).

27. W. B. Tonks, H. J. Melosh, J. Geophys. Res. Planets 98,5319–5333 (1993).

ACKNOWLEDGMENTS

We thank T. Mock for assistance with mass spectrometry.This work was improved by helpful discussions with J. O’Neil,S. Shirey, and B. Wood. We thank S. Shirey, who providedsample VE-32. We also appreciate the helpful comments andsuggestions from three anonymous reviewers. We thankM. Garçon, who developed the four-step 142Nd acquisitionmethod. This work was supported by NSF grant EAR-1265169(Cooperative Studies of the Earth’s Deep Interior program) to

R.J.W. and grant EAR-1250419 to S.M. All data are available inthe main manuscript and supplementary materials.

SUPPLEMENTARY MATERIALS

www.sciencemag.org/content/352/6287/809/suppl/DC1Materials and MethodsFigs. S1 to S6Tables S1 to S5References (28–56)

12 November 2015; accepted 5 April 201610.1126/science.aad8563

SLEEP RESEARCH

Causal evidence for the role of REMsleep theta rhythm in contextualmemory consolidationRichard Boyce,1 Stephen D. Glasgow,2 Sylvain Williams,2*† Antoine Adamantidis2,3*†

Rapid eye movement sleep (REMS) has been linked with spatial and emotional memoryconsolidation. However, establishing direct causality between neural activity during REMSand memory consolidation has proven difficult because of the transient nature of REMSand significant caveats associated with REMS deprivation techniques. In mice, weoptogenetically silenced medial septum g-aminobutyric acid–releasing (MSGABA) neurons,allowing for temporally precise attenuation of the memory-associated theta rhythmduring REMS without disturbing sleeping behavior. REMS-specific optogenetic silencing ofMSGABA neurons selectively during a REMS critical window after learning erasedsubsequent novel object place recognition and impaired fear-conditioned contextualmemory. Silencing MSGABA neurons for similar durations outside REMS episodes had noeffect on memory. These results demonstrate that MSGABA neuronal activity specificallyduring REMS is required for normal memory consolidation.

The physiological function of rapid eyemove-ment sleep (REMS) is unclear (1). Evidencelinking REMS to aspects of memory con-solidation in mammals has been obtainedusing techniques such as statistical corre-

lation, pharmacology, and REMS deprivation(2, 3). However, whether REMS has a directrole in learning and memory remains contro-versial; correlative studies are not definitive,REMS has a transient pattern of occurrencethat prevents REMS-selective pharmacologicalmanipulation, and REMS deprivation has crit-ical caveats that are difficult to fully control for(4, 5).During REMS in mice and rats, a promi-

nent ~7-Hz theta oscillation is observed in lo-cal field potential (LFP) recordings from corticalstructures, including the hippocampus (6, 7).Hippocampal theta rhythms during REMS may

contribute to memory consolidation by pro-viding a mechanism for strengthening placecells formed during prior wakefulness (8, 9).Theta rhythm generation requires an intactmedial septum (MS) (10, 11), although the MSis not involved in REMS generation itself (12, 13).MS g-aminobutyric acid–releasing (MSGABA)neurons project to the hippocampus, prob-ably pacing the hippocampal theta rhythmduring REMS (14–16). In mice, we thereforeused optogenetics to silence MSGABA neuronsand reduce theta activity selectively duringREMS, without disturbing sleeping behavior,to determine whether intact MSGABA neuralactivity during REMS is important for memoryconsolidation.Adeno-associated virus (AAVdj) encoding Arc-

haerhodopsin fused to an enhanced yellow fluo-rescent protein (eYFP) reporter (ArchT-eYFP)was injected into the MS of VGAT::Cre mice (Fig.1A, top). The resulting ArchT-eYFP expressionwas ~95% specific for MSGABA neurons (Fig. 1A,bottom, and fig. S1A), stable, and localized to theMS and the diagonal band of Broca (DBB) regionfor at least 3months after injection.MSGABA neuralprojections were observed throughout the hippo-campus (Fig. 1B).

Whole-cell voltage and current clamp record-ings of ArchT-eYFP–expressing MS neurons inacute brain slices (fig. S1B) revealed hyperpolar-ization (−39.9 ± 6.6 mV) and outward current(293.9 ± 69.2 pA) upon 594-nm light exposure(fig. S1B). Single-unit recordings in behavingtransfected mice (fig. S1C) confirmed that photo-inhibitionduringREMS,non-REMsleep (NREMS),and wakefulness rapidly produced a potent andreversible reduction in spiking of putative MSGABA

neurons (fig. S1D).We next tested the effect of silencing MSGABA

neurons during REMS in freely behaving mice.Photoinhibition with constant light pulses deliv-ered to the MS in ArchT-eYFP–expressing mice(ArchTmice) resulted in significantly (65.3 ± 5.6%)reduced theta power measured from dorsal hippo-campal area CA1 LFP (CA1LFP) recording (Fig. 1D,top). No other frequency bands were affected, andthe spectral profile of the CA1LFP returned tobaseline levels almost immediately upon releaseof MSGABA neurons from photoinhibition (Fig.1D, top, and Fig. 2A). Current source density(CSD) analysis revealed that reduced theta powerwas present in all layers of dorsal hippocampalCA1 (Fig. 2B). Light pulses delivered to the MSof mice only expressing eYFP inMSGABA neurons(YFP control mice) did not affect CA1LFP power(Fig. 1D, bottom), ruling out light as a potentialconfounding factor in these results. Inhibition ofMSGABA neurons did not perturb sleeping behav-ior (Fig. 1D, top), and the probability of statetransition during REMS in ArchT mice was un-altered relative to YFP control mice (n = 30 ArchTmice,n = 19 YFP control mice; P = 0.63, unpairedStudent’s t test).We optogenetically silenced MSGABA neurons

selectively during each REMS episode after ac-quisitionof anovel object place recognition (NOPR)task (Fig. 3A). Mice showed no preference foreither object during initial object habituation[day 1 (D1), task acquisition] (Fig. 3A, right).After acquisition, EEG/CA1LFP/EMG (EEG, elec-troencephalogram; EMG, electromyogram) ac-tivity wasmonitored for 4 hours. Upon entry intoREMS, mice in the ArchT or YFP control grouphad light continuously delivered to the MS untiltransition to another state occurred, at whichtime light delivery ceased until subsequent REMSwas detected (Fig. 3B). A group of ArchT-eYFPexpressing mice that never received light (ArchTcontrol) served as a baseline control for ArchT-eYFP transfection. To determine whether REMSwas a critical factor in our results, a final group

812 13 MAY 2016 • VOL 352 ISSUE 6287 sciencemag.org SCIENCE

1Integrated Program in Neuroscience, McGill University,Montreal, Quebec, Canada. 2Department of Psychiatry, McGillUniversity, Montreal, Quebec, Canada. 3Department ofNeurology and Department of Clinical Research, InselspitalUniversity Hospital, University of Bern, Bern, Switzerland.*Corresponding author. Email: [email protected](S.W.); [email protected] (A.A.) †These authorscontributed equally to this work.

RESEARCH | REPORTSon A

pril 23, 2018

http://science.sciencemag.org/

Dow

nloaded from

on April 23, 2018

http://science.sciencem

ag.org/D

ownloaded from

on A

pril 23, 2018

http://science.sciencemag.org/

Dow

nloaded from

on April 23, 2018

http://science.sciencem

ag.org/D

ownloaded from

of mice expressing ArchT-eYFP (ArchT REMcontrol) underwent the same protocol as micein the ArchT group, with the exception that lightdelivery to the MS was delayed upon detection ofREMS by ~5min (supplementary methods). Afterthe delay, light was delivered continuously to theMS for a duration equaling that of the precedingREMS episode. The result was a pattern of in-hibition that, whereas statistically similar to theREMS-specific pattern that ArchT mice received,occurred primarily when mice were not engagedin REMS. MSGABA neurons were inhibited during91.6 ± 1.3% of the cumulative REMS durationoccurring during the post-test period in ArchTmice (Fig. 3C, top, and fig. S3A), with little inhibi-tion during other states (Fig. 3C, top; laser on 0.2 ±0.0% of NREMS; 1.5 ± 0.2% of wakefulness). Sleeparchitecture was unaffected by REMS-selectiveMSGABA neural photoinhibition during the 4-hourpost-test period (fig. S3, A and B), and the onlysignificant effect on the CA1LFP spectral profile(Fig. 3C, bottom, and figs. S4, A to C, and S5C)was reduced (60.2 ± 2.4%) REMS theta power.Basic firing activity duringwakefulness, NREMS,and REMS of isolated single neurons recordedat the dorsal CA1 cell layer in a subset of mice

during the 4-hour post-test period (supplemen-tary methods) was unaffected by REMS-specificMSGABA neural inhibition (fig. S6, A and B).Analysis of REMS immediately after the 4-hourpost-test period when optogenetic silencing ofMSGABA neurons during REMS had ceased (4- to5-hour post-test/recovery period) revealed animmediate return of REMS CA1LFP spectral pow-er to baseline levels (fig. S4, B and C) and noalteration in cumulative and average episode du-rations in ArchTmice relative to controls (fig. S3C).EEG spindles (tables S1 to S3) and hippocampalripples (tables S6 to S9)measured during NREMSthroughout the course of the NOPR protocol werealso unaltered.On day 2 (D2), object 2 was moved to a new

location while object 1 remained stationary (Fig.3D, left). Because mice preferentially investigatenovel stimuli, if object orientation memory is in-tact,mice should intrinsically investigate object 2more than object 1 (17). Preference for object 2[object 2 discrimination index (DI), supplemen-tary methods] was not different relative to D1testing inArchTmice (Fig. 3D, right, and 3E). Thisresult is not due to the MS light delivery method,because YFP control mice that underwent the

same light delivery protocol (Fig. 3C, top; laser on86.8 ± 4.6% of REMS; 0.2 ± 0.1% of NREMS; 2.7 ±0.3% of wakefulness) had higher D2 object 2 DIrelative to neutral D1 discrimination (Fig. 3D,right, and 3E). ArchT control mice also had in-creased object 2 DI during D2 testing from neu-trality observed on D1 (Fig. 3D, right, and 3E).The co-occurrence of MSGABA neural inhibitionwith REMS was a critical factor in the deficit ob-served in ArchT mice, because ArchT REM con-trol mice demonstrated intact object recognitionmemory comparable to that of YFP control andArchT control mice (Fig. 3D, right, and 3E). Noevidence for differences in locomotion or moti-vation during testing were found between groups(fig. S5, A and B).We next investigated whether normal contex-

tual and emotionalmemory consolidation requiresMSGABA neural activity (Fig. 4). ArchT, YFP con-trol, ArchT control, and ArchT REM control micewere fear-conditioned in a distinct context (con-text A) with three tone-shock events spanning a9.5-min session. Freezing behavior between groupswas not different during conditioning (Fig. 4A).For the subsequent 4 hours, a protocol similar tothat used after D1 NOPR testing was again used

SCIENCE sciencemag.org 13 MAY 2016 • VOL 352 ISSUE 6287 813

Fig. 1. ArchT-mediatedinhibition of MSGABA neuronsduring REMS reduces thetarhythm. (A) Schematic(top left) showing the MSinjection site of Cre-dependentAAV in VGAT::Cre mice. Aftervirus delivery, ArchT-eYFP isinverted in MSGABA neurons,allowing transcription from theEF-1a promoter and subsequentexpression of ArchT-eYFP tooccur in the MS (top right).(Bottom) Cell-specific expres-sion of ArchT-eYFP (green) inMSGABA [glutamic aciddecarboxylase (GAD+)] neurons(magenta) (quantified in fig.S1A). (B) Dense projections inthe hippocampus originatingfrom MSGABA neurons (green).DG, dentate gyrus.(C) Schematic of the in vivorecording configuration; anoptic fiber delivered orangelaser light to the MS, allowingfor optogenetic inhibition ofMSGABA neurons while recordingthe LFP signal from electrodesimplanted in dorsal CA1.(D) Effect of MSGABA neuralinhibition during REMS onCA1LFP and EMG activity. Miceinjected with a control virusresulting in expression of onlyeYFP in MSGABA neuronscontrolled for the use of orangelight (YFP control) [n = 9 forArchTmice, n = 11 for YFP control mice; n.s. = not significant, **P < 0.01, two-way analysis of variance (ANOVA) with Tukey post-hoc test].

60 s

Wake

NREMS

REMS

30

0

0.45

0

0.07

Freq

uenc

y (H

z)

δ θ α β γlow γhighRipple

δ θ α β γlow γhighRipple

0.45 0.07

Arc

hT (

n=9)

YF

P c

ontr

ol (

n=11

)

CA

1 po

wer

(m

V2)

CA1

EMG

Laser on

CA1

EMG

Laser on

0

30

0 Freq

uenc

y (H

z) C

A1

pow

er (

mV

2)

|EM

G| (

mV

2)|E

MG

| (m

V2)

Injection needle

AAVdj-ArchT-eYFP injecton in MS

CA1

DG

DG

Hilus

CA3

Dorsal Hc

CA1 local field potential recording

Optic fiber above MS transfection zone

594 nm light delivery to MS from laser MSGABA photoinhibition

Laser on; light delivery to MS

Baseline Laser on Recovery

Baseline

Laser on

Recovery

1 m

V1

mV

0.5

mV

0.5

mV

1 s

Baseline Laser on Recovery

Baseline

Laser on

Recovery

1 s

0.012

0

Pow

er (mV

2)

0.012

0

Pow

er (mV

2)

100

0

100

0

MS

Injection needle

tract

Scale bars = 10 µm

Scale bar = 200 µm

Scale bar = 100 µm

**

n.s.

n.s.

n.s.

eYFP+ GAD+ Merge

RESEARCH | REPORTS

814 13 MAY 2016 • VOL 352 ISSUE 6287 sciencemag.org SCIENCE

δ(1-4 Hz)

θ(4-10 Hz)

α(10-14 Hz)

β(15-30 Hz)

γ low (30-60 Hz)

Ripple(100-250 Hz)

EMG(30-100 Hz)

CA1RawData

γ high (60-100 Hz)

0.1

mV

0.5 s

MSGABA inhibition

RecoveryBaseline

Frequency (Hz)1.5 20

9.0 E-12

4.0 E-11

4.0 E-10

6.5 E-10

0

0

0

0

CS

D p

ower

(m

V2/

um2)

(R

EM

S)

Oriens

Pyramidale

1

-1

Current source density (m

V/um

2)

Source

Sink

Lacunosum-moleculare

Radiatum

Baseline (REMS) MSGABA inhibition (REMS)

Oriens

Pyramidale

Lacunosum-moleculare

Radiatum

1 mV/um2

0.1 s

50 umspacing

Baseline MSGABA inhibition

Dorsal CA1 probe tract

0.5 mm

Fig. 2. MSGABA neural silencing during REMS reduces theta power in all layers of dorsal hippocampal CA1. (A) Sample raw and filtered traces fromArchT example trace in Fig. 1D, as indicated. Recordings were from the stratum radiatum of dorsal hippocampal CA1. (B) (Left) CSD analysis completed onrecordings obtained from dorsal hippocampal CA1 (top right) using a chronically implanted linear 16-channel probe (50 mm spacing between successivechannels) during REMS under baseline and MSGABA neural inhibition conditions. Graphs at right show layer-specific spectral analysis of CSDs for the baselineversus the MSGABA neural inhibition condition (n = 1 AAVdj-ArchT–injected mouse).

Pro

port

ion

of c

umul

ativ

e ob

ject

ex

plor

atio

n tim

e

1.0

0YFP

control (n=6)

Object 1

Neutral

Pro

port

ion

of c

umul

ativ

e ob

ject

ex

plor

atio

n tim

e

1.0

0

* *

Neutral

Wake

NREMS

REMS

EMG

CA1

Object 2Day 1 NOPR test

Day 2 NOPR test

Frequency (Hz)1

Tes

t ses

sion 1 hr

REMS episode ends; laser off

EMG

CA1

EMG

CA1

5 s

0.012

0

1 mV

0.05 mV0.1 s

0.5 mV0.1 mV

Wake

REMS

NREMS

0-4

hr p

ost-

test

cum

ulat

ive

stat

e pr

opor

tion

Wake

REMS

NREMS

WakeREMS

NREMS

ArchT(n=6)

ArchTcontrol(n=8)

YFP control (n=6)

ArchT(n=6)

ArchTcontrol(n=8)

Novel object place recognition

Return to homecage

RE

M s

leep

CA

1 LF

P

pow

er (

mV

2)

YFP control (n=6) ArchT (n=6)ArchT control (n=8)

Day 1 NOPR post-test protocol schematic

8:30 AM

YFP control

ArchT

ArchT control

8:30 AM - 8:40 AM

Familiarization to objects

Object 1

Object 2

8:30 AM - 8:40 AM

Object 1

Object 2 (new location)

Dorsal Hc

CA1 local field potential recording

Optic fiber above MS transfection zone

594 nm light delivery to MS when laser on

MSGABA

Laser on during REMS; MSGABA neural inhibition

Laser on during REMS; no MSGABA neural inhibition

Laser off at all times; no MSGABA neural inhibition

Day 1 NOPR post-test analysis

-0.2

0.6

0

Day 1 Day 2

Obj

ect 2

dis

crim

inat

ion

inde

x

YFP control (n=6)

ArchT(n=6)

ArchTcontrol(n=8)

Neutral

REMS episode ends; laser off

n.s.

ArchTREM

control(n=12)

ArchTREM

control(n=12)

Object 1 Object 2

*

ArchTREM

control(n=12)

EMG

CA1

ArchT REM control

Wake

NREMS

REMS

Laser on after ~5 min delay following REMS detection; MSGABA neural inhibition for duration of prior

REMS episode occurs during NREMS/wakefulness

Laser on

WakeREMS

NREMS

ArchT REM control (n=12)

20

Fig. 3. Inhibiting MSGABA neuronsselectively during REMS impairssubsequent NOPR. (A) Schematicand object exploration data from D1testing to familiarize mice to object1 and object 2. (B) Mice were returnedto their home cage, and EEG/CA1LFP/EMG activity was monitored for the

subsequent 4 hours. Whenever mice in the ArchT and YFP controlgroups entered into REMS during this period, light was deliveredcontinuously to theMSuntil transition to another state occurred.Micein theArchTcontrol groupdid not have light delivered to theMSat anytime during this period. Mice in the ArchT REM control groupunderwent the same protocol as mice in the ArchT group, with theexception that inhibition was delayed ~5 min to avoid REMS (see textfordetails). (C) (Top)Stateanalysisof4-hourD1post-testperiod (n=6ArchTmice, n = 6 YFPcontrol mice, n = 8ArchTcontrol mice, n = 12ArchT REM control mice; n.s. = not significant, two-way ANOVA).(Bottom) Corresponding spectral analysis of CA1LFP recordings

during REMS. (D) Schematic and object exploration data fromD2 testing; only object 2 placement differed relative to D1. (E) Analysis of object 2 preference betweenD1 and D2 (n = 6 ArchTmice, n = 6 YFP control mice, n = 8 ArchTcontrol mice, n = 12 ArchT REM control mice; *P < 0.05, paired Student’s t test).

RESEARCH | REPORTS

(see Fig. 3B for a detailed schematic). ArchTmicereceived selective MSGABA neural inhibition dur-ing REMS (laser on 92.2 ± 2.2% of REMS; 0.2 ±0.0% of NREMS; 1.6 ± 0.2% of wakefulness) dur-ing the 4-hour post–fear conditioning period(Fig. 4B). No difference in sleep-wake architecture(Fig. 4B, top, and fig. S7, A to C) or NREMS EEGspindle activity (tables S4 and S5) was foundbetween groups, and the only significant effect onthe CA1LFP spectral profile was a 57.8 ± 5.7%reduction in theta power in ArchT mice (Fig. 4B,bottom, and fig. S8, A to C).The next day mice were first tested for con-

textual recall memory (Fig. 4C) followed by cuedrecall memory (Fig. 4D). Mice were placed in con-text A for 10 min, where conditioning had oc-curred the prior day, and allowed to move freelywithout any tone or shock. ArchT mice frozeless than YFP control, ArchT control, and ArchTREM control mice (Fig. 4C, right). One hourlater, mice were placed in a novel context (con-text B) for 9.5 min for cued recall testing, and asequence of tones identical to those from prior

fear conditioning was played. Freezing behav-ior was not different between the ArchT, YFPcontrol, ArchT control, or ArchT REM controlgroups,with each group showing a robust freez-ing response selectively to the cue (tone) (Fig.4D, right).NOPRand fear-conditioned contextualmemory

in these experimentswere probably hippocampus-dependent (18, 19). Considering the potential im-portance of hippocampal REMS theta oscillationsin processing place cell information (8, 9), theimpairments reported here could result fromdisrupted theta-dependent plasticity in hippo-campal neurons during REMS after initial mem-ory consolidation. REMS may also contributeto the homeostasis of network excitability(20, 21). Disruption of hippocampal homeosta-sis could have contributed to the memory im-pairment we observed, although analysis ofCA1 unit data did not reveal any clear indica-tion of altered activity resulting from MSGABA

neural inhibition during REMS. Extrahippocam-pal inputs also must be considered, given known

MS projection patterns (22). Indeed, currentsource density analysis from CA1 during REMSindicated a reduction in theta rhythm powerat all layers upon MSGABA neural inhibition.Thus, in addition to disrupted input from theSchaffer-collaterals, input from the entorhinalcortex via the perforant path was also disrupted.Given the importance of these inputs in spatialmemory and hippocampal place cell activity(18, 23), their disruption may be a mechanisminvolved in the blockade of consolidation weobserved. In summary, our data provide exper-imental proof in a mouse model that MSGABA

neural activity occurring specifically during REMSafter acquisition of a NOPR task or fear condi-tioning is critical for normal spatial and contex-tual memory consolidation.

REFERENCES AND NOTES

1. J. M. Siegel, Nature 437, 1264–1271 (2005).2. S. Diekelmann, J. Born, Sleep 11, 114–126 (2010).3. R. Stickgold, M. P. Walker, Sleep 28, 1225–1227 (2005).4. J. M. Siegel, Science 294, 1058–1063 (2001).

SCIENCE sciencemag.org 13 MAY 2016 • VOL 352 ISSUE 6287 815

15

0

Fre

ezin

g du

ratio

n (s

)

9.50 Elapsed test time (Minutes)

15

0

Fre

ezin

g du

ratio

n (s

)

Ton

e

Ton

e

Ton

e

8:30 AM - 8:40 AM

Day 1 fear conditioning

10

YFP control (n=11) ArchT (n=9)ArchT control (n=9)

7:30 AM - 7:40 AM

Elapsed test time (Minutes)

Day 2 context memory recall test

Context AContext An.s.

15

0

Sho

ck

Sho

ck

Sho

ck

60

0

0

∗†∗∗‡ ∗∗†

∗‡

Unp

roce

ssed

Bin

ned

Unp

roce

ssed

Bin

ned

ArchT REM control (n=5)

YFP control (n=11) ArchT (n=9)ArchT control (n=9)ArchT REM control (n=5)

9.50

Ton

e

Ton

e

Ton

e

Day 2 cued memory recall test

Elapsed test time (Minutes)

8:30 AM - 8:40 AM

Fre

ezin

g du

ratio

n (s

)

Context Bn.s.

30

0

30

0

Unp

roce

ssed

Bin

ned

YFP control (n=11) ArchT (n=9)ArchT control (n=9)ArchT REM control (n=5)

Wake

REMS

NREMS4 hr

pos

t-te

st

cum

ulat

ive

stat

e pr

opor

tion

Wake

REMS

NREMS

WakeREMS

NREMS

Day 1 fear conditioning post-test analysis

n.s.

WakeREMS

NREMS

Laser on

Frequency (Hz)1

0.01

0

RE

M s

leep

CA

1 LF

P

pow

er (

mV

2)

YFP control (n=11) ArchT (n=9)ArchT control (n=9)ArchT REM control (n=5)

20

Fig. 4. Inhibiting MSGABA neurons selectively during REMS after fearconditioning impairs contextual memory. (A) Fear conditioning schematicand freezing data. (B) Immediately after conditioning, mice were returned totheir home cage, where they underwent the same procedure as thatdescribed after D1 NOPR testing (a full schematic is in Fig. 3B). (Top) Stateanalysis of 4-hour post-conditioning period (n = 9 ArchT mice, n = 11 YFPcontrol mice, n = 9 ArchTcontrol mice, n = 5 ArchT REM control mice; n.s. =not significant, two-way ANOVA). (Bottom) Corresponding spectral analysisof CA1LFP recordings during REMS. (C) D2 contextual recall memory test

schematic and freezing analysis. (D) D2 cued recall memory test schematicand freezing analysis. (A), (C), and (D) Freezing versus time graphs (n = 9ArchTmice, n = 11 YFP control mice, n = 9 ArchTcontrol mice, n = 5 ArchTREM control mice; n.s. = not significant, *P < 0.05, **P < 0.01, two-wayrepeated-measures ANOVA with Tukey post-hoc test; †YFP control versusArchTmice, ‡ArchTcontrol versus ArchTmice, #ArchT REM control versusArchT mice). (A) and (D) Statistical results confirmed with Kruskal-Wallistest. Red lines indicate unprocessed data bin boundaries used for statistics(top right plot).

RESEARCH | REPORTS

5. R. P. Vertes, Neuron 44, 135–148 (2004).6. T. E. Robinson, R. C. Kramis, C. H. Vanderwolf, Brain Res. 124,

544–549 (1977).7. G. Buzsáki, Neuron 33, 325–340 (2002).8. G. R. Poe, D. A. Nitz, B. L. McNaughton, C. A. Barnes, Brain

Res. 855, 176–180 (2000).9. K. Louie, M. A. Wilson, Neuron 29, 145–156 (2001).10. J. D. Green, A. A. Arduini, J. Neurophysiol. 17, 533–557

(1954).11. S. J. Mitchell, J. N. Rawlins, O. Steward, D. S. Olton,

J. Neurosci. 2, 292–302 (1982).12. B. E. Jones, Neuroscience 40, 637–656 (1991).13. R. E. Brown, R. Basheer, J. T. McKenna, R. E. Strecker,

R. W. McCarley, Physiol. Rev. 92, 1087–1187 (2012).14. H. Petsche, C. Stumpf, G. Gogolak, Electroencephalogr. Clin.

Neurophysiol. 14, 202–211 (1962).15. A. P. Simon, F. Poindessous-Jazat, P. Dutar, J. Epelbaum,

M. H. Bassant, J. Neurosci. 26, 9038–9046 (2006).16. K. Tóth, T. F. Freund, R. Miles, J. Physiol. 500, 463–474

(1997).

17. M. Antunes, G. Biala, Cogn. Process. 13, 93–110(2012).

18. M. B. Moser, D. C. Rowland, E. I. Moser, Cold Spring HarborPerspect. Biol. 7, 1–15 (2016).

19. J. J. Kim, M. S. Fanselow, Science 256, 675–677(1992).

20. A. A. Borbély, Hum. Neurobiol. 1, 195–204 (1982).21. A. D. Grosmark, K. Mizuseki, E. Pastalkova, K. Diba, G. Buzsáki,

Neuron 75, 1001–1007 (2012).22. R. C. Meibach, A. Siegel, Brain Res. 119, 1–20 (1977).23. M. E. Hasselmo, Eur. J. Neurosci. 28, 1301–1315 (2008).

ACKNOWLEDGMENTS

We thank all members of the Williams and Tidis labs for theirhelpful comments on the manuscript, E. Boyden for viralconstructs, and B. Lowell and L. Vong for mice. R.B. wassupported by an Alexander Graham Bell Canada GraduateScholarship [Natural Sciences and Engineering ResearchCouncil of Canada (NSERC)] while completing this research.

S.D.G. was supported by a postdoctoral fellowship fromFonds de la Recherche en Santé du Québec. S.W. is supportedby the Canadian Institutes of Health Research (CIHR) andNSERC. A.A. is supported by the Human Frontier ScienceProgram (RGY0076/2012), the Douglas Foundation, McGillUniversity, the Canadian Fund for Innovation (CFI), the CanadianResearch Chair (CRC Tier 2), CIHR, NSERC, the Swiss NationalScience Foundation, the Inselspital, and the Universityof Bern. All data are available in the supplementary materials.We declare no conflicts of interest.

SUPPLEMENTARY MATERIALS

www.sciencemag.org/content/352/6287/812/suppl/DC1Materials and MethodsFigs. S1 to S8Tables S1 to S9References (24–27)

9 October 2015; accepted 24 March 201610.1126/science.aad5252

ZIKA VIRUS

Zika virus impairs growth in humanneurospheres and brain organoidsPatricia P. Garcez,2,1* Erick Correia Loiola,1† Rodrigo Madeiro da Costa,1†Luiza M. Higa,3† Pablo Trindade,1† Rodrigo Delvecchio,3

Juliana Minardi Nascimento,1,4 Rodrigo Brindeiro,3

Amilcar Tanuri,3 Stevens K. Rehen1,2*

Since the emergence of Zika virus (ZIKV), reports of microcephaly have increasedconsiderably in Brazil; however, causality between the viral epidemic and malformationsin fetal brains needs further confirmation. We examined the effects of ZIKV infectionin human neural stem cells growing as neurospheres and brain organoids. Usingimmunocytochemistry and electron microscopy, we showed that ZIKV targetshuman brain cells, reducing their viability and growth as neurospheres and brainorganoids. These results suggest that ZIKV abrogates neurogenesis duringhuman brain development.

Primarymicrocephaly is a severe brainmal-formation characterized by the reductionof the head circumference. Patients dis-play a heterogeneous range of brain im-pairments that compromise motor, visual,

hearing, and cognitive functions (1).Microcephaly is associated with decreased

neuronal production as a consequence of pro-liferative defects and death of cortical progenitorcells (2). During pregnancy, the primary etiologyof microcephaly varies from genetic mutationsto external insults. The so-called TORCHS fac-tors (toxoplasmosis, rubella, cytomegalovirus,herpes virus, and syphilis) are the main con-genital infections that compromise brain devel-opment in utero (3).

An increase in the rate of microcephaly inBrazil has been associated with the recent out-break of Zika virus (ZIKV) (4, 5), a flavivirus thatis transmitted by mosquitoes (6) and sexually(7–9). So far, ZIKV has been described in theplacenta and amniotic fluid of microcephalicfetuses (10–13) and in the blood of microcephalicnewborns (11, 14). ZIKV had also been detectedwithin the brain of a microcephalic fetus (13, 14),and recently, direct evidence has emerged thatZIKV is able to infect and cause the death ofneural stem cells (15).We used human induced pluripotent stem

(iPS) cells cultured as neural stem cells (NSCs),neurospheres, and brain organoids to explorethe consequences of ZIKV infection during neu-rogenesis and growth with three-dimensionalculture models. Human iPS-derived NSCs wereexposed to ZIKV [multiplicity of infection (MOI),0.25 to 0.0025]. After 24 hours, ZIKV was de-tected in NSCs (Fig. 1, A to D); viral envelopeprotein was evident in 10.10% (MOI, 0.025) and21.7% (MOI, 0.25) of cells exposed to ZIKV (Fig.1E). Viral RNA was also detected in the super-natant of infected NSCs (MOI, 0.0025) by quan-

titative reverse transcriptase polymerase chainreaction (qRT-PCR) (Fig. 1F), providing evidenceof productive infection.To investigate the effects of ZIKV during

neural differentiation, mock- and ZIKV-infectedNSCs were cultured as neurospheres. After3 days in vitro (DIV), mock-infected NSCs gen-erated round neurospheres. However, ZIKV-infected NSCs generated neurospheres withmorphological abnormalities and cell detach-ment (Fig. 2B). After 6 DIV, hundreds of neu-rospheres grew under mock conditions (Fig. 2,C and E). In ZIKV-infected NSCs (MOI, 2.5 to0.025), only a few neurospheres survived (Fig.2, D and E).Mock-infected neurospheres presented the

expected ultrastructural morphology of the nu-cleus and mitochondria (Fig. 3A). Viral particleswere present in ZIKV-infected neurospheres,similar to those observed in murine glial andneuronal cells (16). ZIKV was bound to the mem-branes and observed in mitochondria and ves-icles of cells within infected neurospheres(arrows in Fig. 3, B and F). Apoptotic nuclei,a hallmark of cell death, were observed in allZIKV-infected neurospheres that we analyzed(Fig. 3B). ZIKV-infected cells in neurospherespresented smooth membrane structures (Fig. 3,B and F), similar to other cell types infectedwith dengue virus (17). These results suggestthat ZIKV induces cell death in human neuralstem cells and thus impairs the formation ofneurospheres.To further investigate the impact of ZIKV

infection during neurogenesis, human iPS-derived brain organoids (18) were exposedto ZIKV and observed for 11 DIV (Fig. 4). Thegrowth rates of 12 individual organoids (sixmock- and six ZIKV-infected) were measuredduring this period (Fig. 4, A to D). As a resultof ZIKV infection, the average growth areaof ZIKV-exposed organoids was reduced by40% compared with brain organoids undermock conditions [0.624 ± 0.064 mm2 for ZIKV-exposed organoids versus 1.051 ± 0.1084 mm2

for mock-infected organoids (normalized);Fig. 4E].

816 13 MAY 2016 • VOL 352 ISSUE 6287 sciencemag.org SCIENCE

1D'Or Institute for Research and Education (IDOR), Rio deJaneiro, Brazil. 2Institute of Biomedical Sciences, FederalUniversity of Rio de Janeiro, Rio de Janeiro, Brazil. 3Instituteof Biology, Federal University of Rio de Janeiro, Rio deJaneiro, Brazil. 4Institute of Biology, State University ofCampinas, Campinas, Brazil.*Corresponding author. Email: [email protected] (P.P.G.);[email protected] (S.K.R.) †These authors contributedequally to this work.

RESEARCH | REPORTS

Causal evidence for the role of REM sleep theta rhythm in contextual memory consolidationRichard Boyce, Stephen D. Glasgow, Sylvain Williams and Antoine Adamantidis

DOI: 10.1126/science.aad5252 (6287), 812-816.352Science 

, this issue p. 812; see also p. 770Sciencenon-REM sleep or wakefulness had no effect on memory.occurred in a critical time window immediately after training. Disrupting the same system for similar durations duringKocsis). Both object recognition memory and contextual fear memory were impaired. This consolidation mechanism

used optogenetics to inhibit theta oscillations in the mouse hippocampus during REM sleep (see the Perspective byet al.The role of REM (rapid eye movement) sleep for memory consolidation has been discussed for a long time. Boyce

Let sleeping mice remember

ARTICLE TOOLS http://science.sciencemag.org/content/352/6287/812

MATERIALSSUPPLEMENTARY http://science.sciencemag.org/content/suppl/2016/05/11/352.6287.812.DC1

CONTENTRELATED

http://stm.sciencemag.org/content/scitransmed/4/129/129ra43.fullhttp://stm.sciencemag.org/content/scitransmed/5/179/179ra44.fullhttp://stm.sciencemag.org/content/scitransmed/5/198/198ra105.fullhttp://stm.sciencemag.org/content/scitransmed/7/305/305ra146.fullhttp://science.sciencemag.org/content/sci/352/6287/770.full

REFERENCES

http://science.sciencemag.org/content/352/6287/812#BIBLThis article cites 26 articles, 4 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.Sciencelicensee American Association for the Advancement of Science. No claim to original U.S. Government Works. The title Science, 1200 New York Avenue NW, Washington, DC 20005. 2017 © The Authors, some rights reserved; exclusive

(print ISSN 0036-8075; online ISSN 1095-9203) is published by the American Association for the Advancement ofScience

on April 23, 2018

http://science.sciencem

ag.org/D

ownloaded from