6
Synapse-specific regulation of AMPA receptor function by PSD-95 Jean-Claude Be ´ı¨que, Da-Ting Lin, Myoung-Goo Kang, Hiro Aizawa, Kogo Takamiya, and Richard L. Huganir* Department of Neuroscience, Johns Hopkins University School of Medicine and Howard Hughes Medical Institute, 725 North Wolfe Street, Baltimore, MD 21205 Contributed by Richard L. Huganir, October 2, 2006 (sent for review September 15, 2006) PSD-95 is a major protein found in virtually all mature excitatory glutamatergic synapses in the brain. Here, we have addressed the role of PSD-95 in controlling glutamatergic synapse function by generating and characterizing a PSD-95 KO mouse. We found that the -amino-3-hydroxy-5-methylisoxazole-4-propionic acid (AMPA) subtype of glutamate receptor (AMPAR)-mediated synaptic trans- mission was reduced in these mice. Two-photon (2P) uncaging of MNI-glutamate onto individual spines suggested that the decrease in AMPAR function in the PSD-95 KO mouse stems from an increase in the proportion of ‘‘silent’’ synapses i.e., synapses containing N-methyl-D-aspartate (NMDA) receptors (NMDARs) but no AM- PARs. Unexpectedly, the silent synapses in the KO mouse were located onto morphologically mature spines. We also observed that a significant population of synapses appeared unaffected by PSD-95 gene deletion, suggesting that the functional role of PSD-95 displays synapse-specificity. In addition, we report that the decay of NMDAR-mediated current was slower in KO mice: The contribution of NR2B subunit containing receptors to the NMDAR- mediated synaptic current was greater in KO mice. The greater occurrence of silent synapses might be related to the greater magnitude of potentiation after long-term potentiation induction observed in these mice. Together, these results suggest a synapse- specific role for PSD-95 in controlling synaptic function that is independent of spine morphology. glutamate receptors hippocampus spines synaptic transmission two-photon uncaging G lutamate, the major excitatory neurotransmitter in the brain, activates ionotropic glutamate receptors of the -amino-3-hydroxy-5-methylisoxazole-4-propionic acid (AMPA), N-methyl-D-aspartate (NMDA), and kainate sub- types. There is considerable interest in elucidating the molecular mechanisms that controls synaptic targeting and trafficking of these receptors, in part because of their role in the induction and expression of various forms of synaptic plasticity (1). These receptors are imbedded in an electron-dense structure, the postsynaptic density (PSD), which is believed to contain key molecules involved in the regulation of glutamate receptor targeting and trafficking. PSD-95, a member of the membrane- associated guanylate kinase (MAGUK) superfamily of proteins, is a core component of the PSD and is thought to be important in the control of excitatory synapse function (2, 3). Because of its interaction with the cytoplasmic domains of NMDA receptor (NMDAR) subunits, it has long been suspected that PSD-95 might control the synaptic targeting of NMDARs (4, 5). How- ever, more recent studies based on sustained or transient over- expression of PSD-95 in neurons have forced a reassessment of this view by suggesting that the primary role of PSD-95 is restricted to controlling AMPAR synaptic expression (6–10). Intriguingly, previous work on a PSD-95 KO mouse, reported no apparent changes in either AMPAR or NMDAR function (11). The interpretation of these data are, however, complicated by the fact that these PSD-95 KO mice still express, albeit at low levels, a functional truncated form of PSD-95 (7, 10). Here, we have reexamined the role of PSD-95 in glutamatergic neurotransmission in the hippocampus by carrying out electro- physiological recordings in a mouse line carrying a complete PSD-95 gene deletion. Our results outline a defect in AMPAR- mediated transmission in the hippocampus of the KO mice. This effect can fully be accounted for by the greater proportion of ‘‘silent’’ synapses found in KO mice as determined by two- photon (2P) uncaging of MNI-glutamate (MNI-GLU). Interest- ingly, the defect in AMPAR function induced by PSD-95 gene deletion appeared to be restricted to only a subpopulation of synapses, thereby suggesting that PSD-95 displays synapse spec- ificity in its actions. In addition, these silent synapses found in KO mice were observed on morphologically mature spines, suggesting that PSD-95 plays a distinct role in controlling glutamatergic synapse function and spine morphology. Results A previous study using gene targeting of the PSD-95 gene resulted in the expression of a truncated form of PSD-95 that contained the N-terminal portion of PSD-95 including PDZ domains 1 and 2 (11). This truncated form has been shown to functionally mimic the effect of full length PSD-95 on AMPAR synaptic expression (7). To generate a complete KO and mini- mize the possibility of generating a functional PSD-95 deletion, a targeting construct was designed to delete PDZ 1 and 2 and generate an out-of-frame PSD-95 transcript (Fig. 5A, which is published as supporting information on the PNAS web site). By using this targeting construct, the PSD-95 gene was disrupted by homologous recombination in embryonic stem cells. Proper targeting was confirmed by Southern blot analysis of genomic DNA using the probe depicted in Fig. 5B and by PCR analysis. Western blot analysis confirmed that PSD-95 protein was not expressed in PSD-95 / (KO) mice (Fig. 5C). Similar results were obtained with several PSD-95 antibodies (data not shown). No gross anatomical or obvious behavioral abnormalities were apparent in these mice. The AMPAR to NMDAR Ratio of Excitatory Postsynaptic Currents Is Reduced in PSD-95 / Mice. To examine the effect of a complete PSD-95 KO on glutamatergic transmission, we first determined Author contributions: J.-C.B. and D.-T.L. contributed equally to this work; J.-C.B., D.-T.L., K.T., and R.L.H. designed research; J.-C.B., D.-T.L., M.-G.K., and H.A. performed research; J.-C.B., D.-T.L., and K.T. contributed new reagentsanalytic tools; J.-C.B., D.-T.L., and M.-G.K. analyzed data; and J.-C.B. and R.L.H. wrote the paper. Conflict of interest statement: Under a licensing agreement between Upstate Group, Inc. and The Johns Hopkins University, R.L.H. is entitled to a share of royalties received by the University on sales of products described in this article. R.L.H. is a paid consultant to Upstate Group, Inc. The terms of this arrangement are being managed by The Johns Hopkins University in accordance with its conflict-of-interest policies. Abbreviations: 2P, two-photon; AMPA, -amino-3-hydroxy-5-methylisoxazole-4-propionic acid; AMPAR, AMPA receptor; DL-AP5, D()-2-amino-5-phosphonovaleric acid; EPSC, ex- citatory postsynaptic current; eEPSC, evoked EPSC; LTP, long-term potentiation; mEPSC, miniature EPSC; MNI-GLU, MNI-Glutamate, 4-methoxy-7-nitroindolinyl-caged-glutamate; NMDA, N-methyl-D-aspartate; NMDAR, NMDA receptor; P(n), postnatal day (n); PSD, post synaptic density. *To whom correspondence should be addressed. E-mail: [email protected]. © 2006 by The National Academy of Sciences of the USA www.pnas.orgcgidoi10.1073pnas.0608492103 PNAS December 19, 2006 vol. 103 no. 51 19535–19540 NEUROSCIENCE Downloaded by guest on June 8, 2021

Synapse-specific regulation of AMPA receptor function by ...PSD-95 gene deletion, suggesting that the functional role of PSD-95 displays synapse-specificity. In addition, we report

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • Synapse-specific regulation of AMPA receptorfunction by PSD-95Jean-Claude Béı̈que, Da-Ting Lin, Myoung-Goo Kang, Hiro Aizawa, Kogo Takamiya, and Richard L. Huganir*

    Department of Neuroscience, Johns Hopkins University School of Medicine and Howard Hughes Medical Institute, 725 North Wolfe Street,Baltimore, MD 21205

    Contributed by Richard L. Huganir, October 2, 2006 (sent for review September 15, 2006)

    PSD-95 is a major protein found in virtually all mature excitatoryglutamatergic synapses in the brain. Here, we have addressed therole of PSD-95 in controlling glutamatergic synapse function bygenerating and characterizing a PSD-95 KO mouse. We found thatthe �-amino-3-hydroxy-5-methylisoxazole-4-propionic acid (AMPA)subtype of glutamate receptor (AMPAR)-mediated synaptic trans-mission was reduced in these mice. Two-photon (2P) uncaging ofMNI-glutamate onto individual spines suggested that the decreasein AMPAR function in the PSD-95 KO mouse stems from an increasein the proportion of ‘‘silent’’ synapses i.e., synapses containingN-methyl-D-aspartate (NMDA) receptors (NMDARs) but no AM-PARs. Unexpectedly, the silent synapses in the KO mouse werelocated onto morphologically mature spines. We also observedthat a significant population of synapses appeared unaffected byPSD-95 gene deletion, suggesting that the functional role ofPSD-95 displays synapse-specificity. In addition, we report that thedecay of NMDAR-mediated current was slower in KO mice: Thecontribution of NR2B subunit containing receptors to the NMDAR-mediated synaptic current was greater in KO mice. The greateroccurrence of silent synapses might be related to the greatermagnitude of potentiation after long-term potentiation inductionobserved in these mice. Together, these results suggest a synapse-specific role for PSD-95 in controlling synaptic function that isindependent of spine morphology.

    glutamate receptors � hippocampus � spines � synaptic transmission �two-photon uncaging

    G lutamate, the major excitatory neurotransmitter in thebrain, activates ionotropic glutamate receptors of the�-amino-3-hydroxy-5-methylisoxazole-4-propionic acid(AMPA), N-methyl-D-aspartate (NMDA), and kainate sub-types. There is considerable interest in elucidating the molecularmechanisms that controls synaptic targeting and trafficking ofthese receptors, in part because of their role in the induction andexpression of various forms of synaptic plasticity (1). Thesereceptors are imbedded in an electron-dense structure, thepostsynaptic density (PSD), which is believed to contain keymolecules involved in the regulation of glutamate receptortargeting and trafficking. PSD-95, a member of the membrane-associated guanylate kinase (MAGUK) superfamily of proteins,is a core component of the PSD and is thought to be importantin the control of excitatory synapse function (2, 3). Because ofits interaction with the cytoplasmic domains of NMDA receptor(NMDAR) subunits, it has long been suspected that PSD-95might control the synaptic targeting of NMDARs (4, 5). How-ever, more recent studies based on sustained or transient over-expression of PSD-95 in neurons have forced a reassessment ofthis view by suggesting that the primary role of PSD-95 isrestricted to controlling AMPAR synaptic expression (6–10).Intriguingly, previous work on a PSD-95 KO mouse, reported noapparent changes in either AMPAR or NMDAR function (11).The interpretation of these data are, however, complicated bythe fact that these PSD-95 KO mice still express, albeit at lowlevels, a functional truncated form of PSD-95 (7, 10).

    Here, we have reexamined the role of PSD-95 in glutamatergicneurotransmission in the hippocampus by carrying out electro-physiological recordings in a mouse line carrying a completePSD-95 gene deletion. Our results outline a defect in AMPAR-mediated transmission in the hippocampus of the KO mice. Thiseffect can fully be accounted for by the greater proportion of‘‘silent’’ synapses found in KO mice as determined by two-photon (2P) uncaging of MNI-glutamate (MNI-GLU). Interest-ingly, the defect in AMPAR function induced by PSD-95 genedeletion appeared to be restricted to only a subpopulation ofsynapses, thereby suggesting that PSD-95 displays synapse spec-ificity in its actions. In addition, these silent synapses found inKO mice were observed on morphologically mature spines,suggesting that PSD-95 plays a distinct role in controllingglutamatergic synapse function and spine morphology.

    ResultsA previous study using gene targeting of the PSD-95 generesulted in the expression of a truncated form of PSD-95 thatcontained the N-terminal portion of PSD-95 including PDZdomains 1 and 2 (11). This truncated form has been shown tofunctionally mimic the effect of full length PSD-95 on AMPARsynaptic expression (7). To generate a complete KO and mini-mize the possibility of generating a functional PSD-95 deletion,a targeting construct was designed to delete PDZ 1 and 2 andgenerate an out-of-frame PSD-95 transcript (Fig. 5A, which ispublished as supporting information on the PNAS web site). Byusing this targeting construct, the PSD-95 gene was disrupted byhomologous recombination in embryonic stem cells. Propertargeting was confirmed by Southern blot analysis of genomicDNA using the probe depicted in Fig. 5B and by PCR analysis.Western blot analysis confirmed that PSD-95 protein was notexpressed in PSD-95�/�(KO) mice (Fig. 5C). Similar results wereobtained with several PSD-95 antibodies (data not shown). Nogross anatomical or obvious behavioral abnormalities wereapparent in these mice.

    The AMPAR to NMDAR Ratio of Excitatory Postsynaptic Currents IsReduced in PSD-95�/� Mice. To examine the effect of a completePSD-95 KO on glutamatergic transmission, we first determined

    Author contributions: J.-C.B. and D.-T.L. contributed equally to this work; J.-C.B., D.-T.L.,K.T., and R.L.H. designed research; J.-C.B., D.-T.L., M.-G.K., and H.A. performed research;J.-C.B., D.-T.L., and K.T. contributed new reagents�analytic tools; J.-C.B., D.-T.L., and M.-G.K.analyzed data; and J.-C.B. and R.L.H. wrote the paper.

    Conflict of interest statement: Under a licensing agreement between Upstate Group, Inc.and The Johns Hopkins University, R.L.H. is entitled to a share of royalties received by theUniversity on sales of products described in this article. R.L.H. is a paid consultant to UpstateGroup, Inc. The terms of this arrangement are being managed by The Johns HopkinsUniversity in accordance with its conflict-of-interest policies.

    Abbreviations: 2P, two-photon; AMPA, �-amino-3-hydroxy-5-methylisoxazole-4-propionicacid; AMPAR, AMPA receptor; DL-AP5, D(�)-2-amino-5-phosphonovaleric acid; EPSC, ex-citatory postsynaptic current; eEPSC, evoked EPSC; LTP, long-term potentiation; mEPSC,miniature EPSC; MNI-GLU, MNI-Glutamate, 4-methoxy-7-nitroindolinyl-caged-glutamate;NMDA, N-methyl-D-aspartate; NMDAR, NMDA receptor; P(n), postnatal day (n); PSD, postsynaptic density.

    *To whom correspondence should be addressed. E-mail: [email protected].

    © 2006 by The National Academy of Sciences of the USA

    www.pnas.org�cgi�doi�10.1073�pnas.0608492103 PNAS � December 19, 2006 � vol. 103 � no. 51 � 19535–19540

    NEU

    ROSC

    IEN

    CE

    Dow

    nloa

    ded

    by g

    uest

    on

    June

    8, 2

    021

  • the ratio of AMPA to NMDA receptor components of evokedexcitatory postsynaptic currents (eEPSCs) in pyramidal neuronsof the CA1 region of the hippocampus. We recorded eEPSCswhile clamping the cell at �40 mV. In these conditions, bothAMPA and NMDARs are activated by synaptically releasedglutamate, and their respective contribution to the eEPSC wasdetermined pharmacologically by administration of theNMDAR antagonist DL-AP5 (100 �M; Fig. 1A1; and seeSupporting Methods, which is published as supporting informa-tion on the PNAS web site) (9, 12, 13). The AMPAR�NMDARratio was significantly reduced in KO compared with age-matched littermate PSD-95�/�mice (WT; Fig. 1 A1).

    Because AMPAR- and NMDAR-mediated synaptic responsesare kinetically distinguishable, we could also approximate AMPARand NMDAR components by measuring the amplitude of thesynaptic current at different time points of the eEPSCs at �40 mV(see Fig. 1A2 and Supporting Methods). Although less precise, thismethod allowed sampling over a much broader population ofrecordings and allowed us to estimate the developmental profile ofthe defect in synaptic transmission induced by genetic deletionof PSD-95 (Fig. 1B). In young animals [postnatal day (P)9–P12], theAMPA�NMDA ratio was not altered in the KO compared withWT mice. However, this ratio was reduced by �30% in KO slicesin a slightly older age group (P14–P20), and this reduction persistedin the older animals tested (Fig. 1B).

    In principle, a change in rectification properties of AMPARscould account for the decrease in AMPAR�NMDAR ratio ob-served at �40 mV even though spermine was not included in ourpipette solution. To control for this possibility, we obtained full I-Vcurves of AMPAR-mediated EPSCs and found no changes inrectification of AMPAR currents between WT and KO mice (Fig.6, which is published as supporting information on the PNAS website). Together, these results outline a general defect in glutama-tergic transmission in mice lacking PSD-95. In addition, the devel-opmental profile of this defect is broadly consistent with the knownexpression profile of PSD-95, which begins to be highly expressedin the hippocampus at around P10 (14).

    Frequency, but Not Amplitude, of AMPAR-Mediated Miniature EPSCs(mESPCs) Is Reduced in PSD-95�/� Mice. A reduction in the AMPA�NMDA ratio could reflect a reduction in the number (or function)

    of AMPARs, an increase in the number (or function) of NMDAR,or a combination of both. To distinguish between these possibilities,we next recorded AMPAR-mediated mEPSCs. During the devel-opmental period studied here (P8–P25), we observed in WT micean increase in the frequency of mEPSCs with increasing age (Fig.1C). This developmentally regulated increase in mEPSC frequencywas largely abolished in KO mice. Thus, by the 3rd week of life, thefrequency of mEPSCs was reduced by �50% in KO mice (Fig. 1C).We observed no differences in the amplitude of mEPSCs betweenWT and KO mice at all ages tested (Fig. 1C).

    A decrease in the frequency of mEPSCs can reflect either adecrease in the probability of neurotransmitter release or adecrease in the number of AMPAR-containing synapses. Todistinguish between these possibilities, we carried out paired-pulse ratio analysis, a measure of presynaptic neurotransmitterrelease probability, and found no difference between WT andKO mice (Fig. 7, which is published as supporting information onthe PNAS web site). The reduction in frequency of mEPSCs inKO mice may thus reflect a reduction in the number of AMPAR-containing synapses. Furthermore, because this reduction inmEPSC frequency can quantitatively account for the reductionin AMPAR�NMDAR ratio observed in KO mice, these resultssuggest that synaptic NMDAR function is largely unaffected inthese mice. Together, these observations suggest the intriguingpossibility that a greater proportion of synapses in KO micewould be functionally silent [i.e., containing NMDARs but noAMPARs (15, 16)].

    2P Uncaging of MNI-GLU Reveals Silent Synapses in Developing CA1Pyramidal Neurons. The existence of silent synapses was originallydemonstrated by minimal stimulation experiments, which arebelieved to monitor synaptic responses elicited by activation ofsingle synapses (16, 17). These experiments are laborious and donot lend themselves with ease to quantitative comparison be-tween groups. We therefore systematically probed the functionof individual synapses (i.e., determined whether they are func-tional or silent) by locally and transiently ‘‘uncaging’’ glutamateby laser photolysis of a glutamate analog, MNI-caged-L-glutamate (MNI-GLU). Neurons (filled for �20 min with AlexaFluor 594) were first imaged by confocal microscopy to identifydendritic spines (Fig. 2A). In agreement with a report using

    10 pA100 ms

    Wild type PSD-95 -/-

    P8-10

    P8-10

    P21-P25

    1.2

    1.0

    0.8

    0.6

    0.4

    0.2

    0.0

    Fre

    quen

    cy (

    Hz)

    Am

    plitu

    de (

    pA)

    15

    10

    P14-P20

    P14-P20

    P21-P25

    Cum

    ulat

    ive

    frac

    tion

    P8-10

    0 20 40Amplitude (pA)

    60 800.0

    0.5

    1.0P16-P20

    Amplitude (pA)0 20 40 60 80 100

    0.0

    0.5

    1.0P21-P25

    Amplitude (pA)0 20 40 60 80 100

    0.0

    0.5

    1.0

    AM

    PA

    /NM

    DA

    RA

    TIO

    + 40 mV

    - 60 mV

    1 2

    3 X Tau10 pA10 ms

    Ctl

    D-APV

    NMDA component

    AM

    PA

    /NM

    DA

    RA

    TIO

    1.75

    1.50

    1.25

    1.00

    0.75

    0.50

    0.25

    0

    1.75

    1.50

    1.25

    1.00

    0.75

    0.50

    0.25

    0

    A1 CA2

    AM

    PA

    /NM

    DA

    RA

    TIO

    1.75

    1.50

    1.25

    1.00

    0.75

    0.50

    0.25

    0P9-12

    WTKO

    P21-P24P14-P20

    B

    PSD-95 -/-Wild type

    Fig. 1. AMPAR-function is de-creased in PSD-95�/� mice. (A) In A1,the AMPA�NMDA ratio was deter-mined by subtracting the traces ob-tained in 100 �M D-L AP5 from thoseobtained in its absence (C � 1) andwas found to be significantly lower(P � 0.05, Student’s t test) in KO (n �9) compared with WT (n � 5) mice.These results were obtained in age-matched littermates (P16–P20). InA2, the AMPA�NMDA ratio was cal-culated by estimating the respectiveAMPA (1) and NMDAR (2) current onthe traces at �40 mV based on theirdifferent time courses (as depicted inthe Inset). The AMPA�NMDA ratiowas significantly lower (P � 0.01, Stu-dent’s t test) inKO(n�19)thaninWTmice (n � 18). (B) The AMPA�NMDAratios obtained by the method de-picted in A2 are shown binned ac-cording to the age of the animals(P9–P12, P � 0.07, n � 5 each; P14–20, P � 0.05; n � 7 each; P21–P24, P �0.05; n � 6 for WT and n � 7 for KO). (C) Frequency and amplitude of mEPSCs were binned according to the age of the animals. For the frequency of events: P8–P10,P � 0.3; n � 11 for WT and n � 4 for KO mice; P14–P20, P � 0.05; n � 10 for WT and n � 16 for KO mice; P21–P25, P � 0.05; n � 10 for WT and n � 8 for KO mice. Forthis and subsequent figures, an asterisk indicates statistical significance.

    19536 � www.pnas.org�cgi�doi�10.1073�pnas.0608492103 Béı̈que et al.

    Dow

    nloa

    ded

    by g

    uest

    on

    June

    8, 2

    021

  • similar methodologies (18), bath administration of MNI-GLU (5mM) did not activate any obvious membrane conductance at�70 mV (data not shown). Under these conditions, brief focallaser illumination (720 nm) of a dendritic spine elicited rapidlyactivating and inactivating inward currents (2P-EPSC; n � 100;Fig. 2). The kinetic properties of 2P-EPSCs were indistinguish-able from those displayed by mEPSCs recorded from the samecell (rise, 3.4 � 0.7 ms and 2.7 � 0.5 ms; decay, 8.4 � 1.9 ms and7.8 � 1.5 ms for 2P-EPSCs and mEPSCs, respectively; n � 6; Fig.2B). To determine the spatial profile of the uncaging spot, wefocused the laser beam at fixed spatial intervals (0.5 �m) alongan axis perpendicular to the length of a spine (Fig. 2C) whilerecording the elicited current. The FWHM values of the result-ing Gaussian function (amplitude vs. location) was 1.39 � 0.9�m, whereas that of the fluorescence profile of the spine per sewas 0.65 � 0.01 �m, thereby confirming the local nature of theuncaging event within the constraints expected from the size ofthe diffraction-limited spot in our conditions and from diffusionof uncaged glutamate. To minimize spillover of the uncagingevent, we carefully chose spines that were separated by ��2 �mfrom their closest neighbor. Uncaging MNI-GLU while holdingthe voltage of the neuron at different values yielded currents thatbehaved like those induced by endogenously released glutamate:the currents reversed at �0 mV and displayed slower decay atdepolarized potentials, as expected from the gradual recruit-ment of voltage-dependent NMDARs (Fig. 2D). The currentinduced by uncaging MNI-GLU was mediated by both AMPARsand NMDARs at �40 mV and solely by AMPARs at �70 mV(Fig. 2E). Together, these findings show that uncaging MNI-GLU onto spines activates ionotropic glutamate receptors in amanner closely mimicking activation by endogenous glutamate.

    To determine whether this approach was able to detect silentsynapses, we focused our attention on young (P7–P8) developing rathippocampal slices where silent synapses have been shown to beprevalent (15, 16). We initially uncaged MNI-GLU onto charac-teristically long, thin filopodia-like spines because they are believedto be largely devoid of AMPARs (18). Fig. 2F depicts two suchspines from which we recorded 2P-EPSCs only when the cell wasvoltage-clamped at �40 mV. These filopodia thus appeared toexpress on their surface functional NMDARs but no AMPARsand, as such, fulfilled the definition generally agreed upon for silentsynapses. As reported in ref. 18, we also observed a correlationbetween the amplitude of AMPAR-mediated 2P-EPSC and thevolume of spine heads, but not spine length, at this young age (see

    Fig. 8A, which is published as supporting information on the PNASweb site).

    Altogether, these results indicate that this approach affordsremarkably precise spatial and temporal control of ionotropicglutamate receptor activation at individual spines while alsoallowing correlative analysis of function and morphology.

    There Are More Silent Synapses in PSD-95�/� Mice as Detected by2P-Uncaging of MNI-GLU. At silent synapses, NMDAR-mediatedsynaptic responses are �5–15 pA in amplitude when recorded at�40 mV (16). To detect silent synapses in the most physiologicallyrelevant manner, we adjusted the uncaging laser power to obtainNMDAR-mediated 2P-EPSCs of 5–15 pA at �40 mV. We thenmeasured the AMPAR-mediated 2P-EPSCs at �70 mV andcomputed an AMPA�NMDA ratio for each spine (see SupportingMethods for more details). Examples of spines and corresponding2P-EPSCs traces at �40 and �70 mV are shown in Fig. 3A. Usingthis approach in WT (P13–P16) mice, we found that individualspines exhibited remarkable heterogeneity as inferred from analysisof their AMPAR�NMDAR ratios in response to 2P-uncaging(range from 0.15 to �3; Fig. 3B). The high ratios observed areunlikely to reflect poor voltage clamping of the spine under study(which would lead to an underestimate of NMDAR current at �40mV) because fast AMPAR-mediated 2P-EPSCs can readily be seenon the traces at �40 mV (Fig. 3A). This broad range of AMPA�NMDA ratios for individual spines was also found in neurons fromage-matched KO mice. However, the distribution of those ratioswas highly skewed such that a greater proportion of spines exhibitedvery low, or zero, AMPA�NMDA ratios in KO compared with WTmice (Fig. 3B). We thus readily detected silent synapses only in KOmice. Consistent with our previous observations, the averagedAMPA�NMDA ratios across all spines we analyzed was signifi-cantly lower in the KO compared with WT mice (Fig. 3 B and D).Thus, these results suggest that the reduction in AMPA�NMDAratio observed in KO mice selectively stems from a greater occur-rence of silent synapses. Importantly, the distribution of AMPA�NMDA ratios revealed that some spines in the KO mice exhibitedratios as high as the highest encountered in the WT mice. Theseresults indicate that a subpopulation of synapses appeared unaf-fected by PSD-95 deletion.

    Divergence Between the Effects of PSD-95�/� Gene Deletion onSynaptic Function and Spine Volume. As determined by 2P-uncaging on synapses from young (P7–P8) rat, we found that

    10 msec

    2P-EPSCmEPSC

    E F

    10 msec20 pA

    -60

    mV

    pA

    -40 -20 20 40

    -80

    -60

    -40

    20

    40

    + 40 + 40 + 40

    - 70 - 70 - 7010 pA

    NBQX (20 µM) NBQX (20 µM)and APV (100 µM)

    5 pA

    Position (µm)-3 -2 -1 0 1 2 3

    Am

    plitu

    de (

    pA)

    0

    5

    10

    Max rise slope

    (pA/m

    s)0

    5

    10

    15

    20

    25

    -2 -1 0 1 2A B C D

    20 ms 25 ms

    Fig. 2. The electrophysiological re-sponse to 2P uncaging of MNI-GLUclosely mimics that of endogenouslyreleased glutamate. (A) 2P image of aCA1 pyramidal neuron filled with 40�M Alexa Fluor 594. (B) On this imageof a spine, the yellow dot (Upper)represents the spot of laser illumina-tion. (Lower) An averaged currenttrace depicting a 2P-EPSC induced bylaser illumination is superimposedwith an averaged mEPSC trace col-lected from the same neuron. (C) Thelaser spot was delivered at fixed spa-tial intervals (0.5 �m) around thespine and the amplitude (F; redtrace) and rise time (E) of the 2P-EPSCs are plotted as a function ofdistance from the center of the spine.The intensity profile of the spine isdepicted by the blue trace. (D) Traces depicting 2P-EPSCs while holding the cell at different voltages are shown. The peak amplitude of the 2P-EPSC is plottedas a function of the holding voltage. (E) Traces depicting 2P-EPSCs at �70 mV and �40 mV, elicited from the spine shown, are shown during baseline, after bathadministration of NBQX (20 �M) and after the subsequent administration of DL-APV (100 �M). (F) On these two neighboring spines, displaying a characteristicthin profile, derived from a young (P8) rat, 2P uncaging elicited responses at �40 mV but not at �70 mV.

    Béı̈que et al. PNAS � December 19, 2006 � vol. 103 � no. 51 � 19537

    NEU

    ROSC

    IEN

    CE

    Dow

    nloa

    ded

    by g

    uest

    on

    June

    8, 2

    021

  • small, filopodia-like spines appeared to be largely devoid ofAMPARs (Fig. 2F and ref. 18), although they expressedNMDAR-mediated dependent current (Fig. 2F). We thus hy-pothesized that the silent synapses observed in KO mice wouldpredominantly be located on smaller, less developed, filopodia-like spines. To address this possibility, we plotted the AMPA�NMDA ratio obtained by 2P-uncaging for each spine againsttheir respective volume (see Supporting Methods). In WT mice,we found that spine volume was not correlated with AMPA�NMDA ratios (Fig. 3C; R2 � 0.04, n � 20). This lack ofcorrelation was expected because these recordings were carriedout in mice at an age (P13–P16) where AMPAR current was notcorrelated to spine volume (data not shown) and where silentsynapses were not readily detected under our conditions (Fig.3B). Interestingly, we also found that spine volume was notcorrelated with AMPA�NMDA ratio in the KO mice (Fig. 3C;R2 � 0.08; n � 25). Closer examination revealed that the volumeof spines that were found to be silent in KO mice spanned closeto the entire range of spine volumes of the population. Thus,these results show that the silent synapses observed in KO miceare not restricted to smaller, less ‘‘developed’’ spines.

    Although the results outlined above show that AMPA�NMDAratios were not correlated to spine volume, it remains possible thatthere was a generalized reduction in spine volume in KO mice. Forthe spines from which we computed 2P-AMPA�NMDA ratios, wedid not observe any significant difference in spine volume betweenWT and KO mice (Fig. 3 C and D). To avoid a possible selectionbias toward a morphologically homogenous population, we alsodetermined spine volume from a much broader population ofspines and also found that spine volumes were unchanged in KOmice (Fig. 3E). These results suggest that the general deficit insynaptic function observed after PSD-95 gene deletion is notaccompanied by any significant changes in the average spinevolume. However, we observed a small but significant increase inspine length in the KO mice (see Fig. 9, which is published assupporting information on the PNAS web site).

    Decay Kinetics of NMDAR-Mediated Synaptic Currents Are Slower inthe PSD-95�/� Mice. During the course of the experiments aimedat determining AMPA�NMDA ratio of eEPSCs at �40 mV (Fig.

    1 B and C), we noticed that the late portion of synaptic currentsrecorded at �40 mV (which predominantly reflects activation ofNMDARs) tended to exhibit slower decay in the KO mice (timeconstants: WT, 78.2 � 8.9 ms, n � 17; KO, 106.8 � 7.1 ms, n �15; P � 0.05; current decay fitted with a simple monoexponen-tial). To address this issue more rigorously, we recorded isolatedNMDAR-mediated synaptic currents (in 0.1 mM Mg2��10 mMglycine�10 �M NBQX; Fig. 4A). The isolated NMDAR synapticcurrents in both WT and KO mice were best fitted by a doubleexponential decay (Fig. 10A, which is published as supportinginformation on the PNAS web site). As shown in Fig. 4B, theweighed time constant for NMDAR decay kinetics were signif-icantly longer in KO compared with WT mice.

    There are several mechanisms that could account for thelonger NMDAR decay kinetics observed in KO mice. Studies inheterologous cells have shown that NMDA receptors containingNR2B-D subunits exhibit slower deactivation than those con-taining NR2A (19), raising the possibility that the longer decaykinetics observed in KO mice may reflect synaptic NMDAR ofdifferent subunit composition. The specific changes in decaykinetics of NMDAR-mediated currents (Fig. 10) are fully con-sistent with this idea (19). To further test this possibility, wedetermined the effects of the preferential NR2B-containingNMDAR antagonist ifenprodil on isolated NMDAR-mediatedcurrents. We found that the inhibitory effect of ifenprodil (3�M) on NMDAR-mediated currents was significantly greater inthe KO mice, thereby suggesting that there is a greater contri-bution of NR2B subunit-containing NMDARs to the NMDAR-mediated synaptic current after PSD-95 gene deletion (Fig. 4C).

    Although our 2P-uncaging experiments were initially designed todetermine AMPA�NMDA ratio at individual spines, we could getan estimate of NMDAR kinetics at single spines by fitting to amonoexponential decay of the NMDA portion of the 2P-EPSC at�40 mV. Consistent with our previous observations, we also foundthat the decay kinetics of 2P-NMDAR currents were significantlylonger in the KO compared with WT mice (Fig. 4 D and E).

    There Was No Correlation Between NMDAR Kinetics and AMPARContent at Individual Synapses. It has been well described thatthere is a gradual shift from mainly NR2B- to NR2A-containing

    B1 B2

    C

    AMPA/NMDA ratio(NMDA=5-15 pA)

    AMPA/NMDA ratio(NMDA=5-15 pA)

    AMPA/NMDA ratio(NMDA=5-15 pA)

    PSD-95 -/-Wild type

    0.0 0.5 1.0 1.5 2.0 2.5 3.00

    1

    2

    3

    4

    5

    6

    7

    Cou

    nt

    0.0 0.5 1.0 1.5 2.0 2.5 3.00

    1

    2

    3

    4

    5

    6

    7

    Cou

    nt

    PSD-95 -/-Wild type

    0 0.5 1.0 1.5 2.0 2.5 3.53.00.0

    0.5

    1.0

    Cum

    ulat

    ive

    frac

    tion

    A

    PSD-95 -/-

    Wild type

    10 pA25 ms

    Spine volume (µm3) Spine volume (µm3)

    PSD-95 -/-Wild type

    AM

    PA

    /NM

    DA

    rat

    io(N

    MD

    A=

    5-15

    pA

    )

    0.0 0.1 0.2 0.3 0.40

    1

    2

    3

    4

    0.0 0.1 0.2 0.3 0.40

    1

    2

    3

    4 D

    *

    AM

    PA

    /NM

    DA

    rat

    io(N

    MD

    A=

    5-15

    pA

    )

    0.0 0.30.1 0.2

    0.5

    0.0

    1.0

    1.5

    PSD-95 -/-

    Wild type

    Spi

    ne v

    olum

    e ( µ

    m3 )

    0.0

    0.1

    0.2

    -/-+/+

    E

    Spine volume (µm3)

    Fig. 3. Analysis of glutamatergic synapses by 2P uncaging of MNI-GLU and analysis of spine morphology in PSD-95�/� mice. (A) Images of individual spines fromWT and KO mice are shown with their respective 2P-EPSCs elicited at �70 and �40 mV. (B1) Distributions of 2P-AMPA�NMDA ratio onto individual spines areplotted for both WT and KO mice. (B2) Cumulative distribution plots of the ratios shown in B1. (C) The 2P-AMPAR�NMDAR ratio obtained from uncaging ontoindividual spines is plotted against the volume of those spines. (D) The average AMPAR�NMDAR ratio (WT, 1.13 � 0.14, n � 20; KO, 0.63 � 0.14, n � 26; P � 0.05)obtained for individual spines by 2P-uncaging is plotted on the y axis against the average volume of the spines (WT, 0.16 � 0.02 �m3, n � 20; KO, 0.17 � 0.02�m3, n � 25; P � 0.7) onto which the AMPAR�NMDAR were determined. (E) The volumes of spines were determined from a much broader population than inD; WT, 0.17 � 0.01 �m3, n � 271 spines, 7 cells, 4 mice; KO, 0.17 � 0.01 �m3, n � 509 spines, 12 neurons, 5 mice).

    19538 � www.pnas.org�cgi�doi�10.1073�pnas.0608492103 Béı̈que et al.

    Dow

    nloa

    ded

    by g

    uest

    on

    June

    8, 2

    021

  • NMDARs during postnatal development (20, 21). In addition,the process of synaptic maturation is accompanied by the gradualincorporation of AMPARs into synapses (22). These studiesraised the possibility that the incorporation of AMPARs insynapses would parallel the shift in NMDAR subunit composi-tion from mainly NR2B-containing receptors toward NR2A-containing receptors and that these processes might be causallyrelated. If this were the case, one would expect a tight correlationbetween the NR2 subunit composition of a given synapse and itsAMPA receptor content. To address this issue, we plotted theAMPA�NMDA ratio of 2P-EPSCs against the decay kinetics of2P-NMDAR-mediated responses at �40 mV for all spinesanalyzed (same as those in Fig. 3). Surprisingly, we found nocorrelation between AMPA�NMDAR ratio of 2P-EPSCs and2P-NMDAR-mediated decay kinetics in both WT (R2 � 0.001;n � 20) and in KO mice (R2 � 0.096; n � 26). These resultssuggest that the level of AMPAR expression and the subunit

    composition of NMDAR at individual synapses are not causallyrelated. We also found no correlation between 2P-NMDAR-mediated decay kinetics and spine volume during this develop-mental period in both WT and KO mice (see Fig. 10B).

    The Magnitude of Long-Term Potentiation (LTP) Is Greater in PSD-95�/� Mice. To examine the functional consequences of thecomplete PSD-95 KO on synaptic plasticity, we examined LTPin the CA1 region of the hippocampus using field potentialrecordings in the stratum radiatum. The magnitude of LTP,induced by delivering a � burst-like protocol, was much greaterin KO mice than in WT littermates (field EPSP slope, comparedwith baseline, at 30–40 min after LTP induction: WT, 182 �13%; KO, 331 � 41%; Fig. 11, which is published as supportinginformation on the PNAS web site). This finding is similar to thatreported for mice expressing the truncated form of PSD-95 (11).

    DiscussionPSD-95 and other members of the MAGUK superfamily of pro-teins are thought to be critical for the proper formation andmaintenance of excitatory synapses. Using a combination of genetargeting, electrophysiological, and imaging approaches, we reportthat AMPAR function was impaired in PSD-95 KO mice. Specif-ically, this reduction was caused by a greater proportion of synapsesthat lack functional AMPARs (i.e., silent synapses) rather than bya general decrease of functional AMPAR across all synapses. Thisfinding suggests a synapse selectivity for PSD-95’s action. Ourresults further showed that silent synapses were found on morpho-logically mature spines in the KO mice. These results thus outlinea distinct role of PSD-95 in controlling glutamatergic synapsefunction independent of spine morphology. Moreover, in confir-mation of a previous report, we find that disruption of the PSD-95gene results in a greater magnitude of LTP.

    Complete genetic deletion of PSD-95 leads, by approximatelythe 2nd week of life, to a reduction in the synaptic expression ofAMPARs, but not NMDARs, as was apparent mainly from theobservation that the reduction in the frequency of AMPAR-mediated mEPSCs in KO mice could quantitatively account forthe reduction in AMPA�NMDA ratio. This finding obtained byusing a loss-of-function strategy is conceptually consistent withstudies showing that overexpression of PSD-95 increasesAMPAR function (6, 7, 9, 10). Interestingly, our 2P-uncagingexperiments directly demonstrated a greater proportion of syn-apses that expressed NMDARs, and not AMPARs in the KOmice. These synapses are referred to as silent because NMDARsare nonfunctional at rest. The greater proportion of silentsynapses in KO mice bears important consequences for synapticplasticity because it is believed that silent synapses represent thepreferential site of AMPAR insertion during LTP (23) (but seerefs. 24 and 25). It is thus conceivable that the enhancedmagnitude of LTP we observed in KO mice directly stems fromthe greater proportion of silent synapses. The corollary impli-cation of these findings is that, whereas PSD-95 appears to be akey molecule in controlling synaptic AMPARs expression, itdoes not appear to play an important role in the actual recruit-ment process of AMPARs to synapses during LTP.

    In principle, the reduction in synaptic AMPAR function in KOmice could be caused by a decrease in synaptic expression ofAMPARs across all synapses. Our results, however, support analternate scenario whereby the deficit in synaptic AMPAR expres-sion appears to be restricted to only a subset of synapses. Twoindependent sets of experiments lead to this interpretation. First,the amplitude of mEPSCs was unaltered by PSD-95 gene deletion(there were simply fewer events). Second, the distribution ofAMPA�NMDA ratios obtained from 2P-uncaging onto individualspines revealed that a subset of spines in KO mice displayed equallyhigh ratios to those obtained from spines in WT mice. Together,these results suggest that a population of synapses have matured (at

    Fig. 4. Analysis of NMDAR function in PSD-95�/� mice. (A) Current tracesshowing NMDAR-mediated synaptic currents recorded in 10 �M glycine, 20�M NBQX, and 0.1 mM Mg2�at �60 mV. (B) The decay kinetics of NMDAR-mediated synaptic currents were longer in KO mice (weighted time constants:WT, 75 � 2 ms, n � 8; KO, 113 � 7 ms, n � 8; P � 0.01). (C) The inhibition ofNMDAR-mediated synaptic currents induced by ifenprodil (3 �M) was largerin KO mice (WT, n � 8; KO, n � 9; P � 0.05, unpaired Student’s t test testedbetween 30 and 35 min). Stimulation frequency was 0.06 Hz. (D) Current tracesshowing mixed AMPA and NMDA currents elicited by 2P-uncaging of MNI-GLU recorded in normal Ringer solution while holding the cell at �40 mV inWT and KO mice. (E) The decay kinetics of the NMDAR portion of the mixed2P-current was approximated by fitting a single exponential decay and waslonger in KO mice (WT, 173 � 13 ms, n � 20; KO, 250 � 16 ms; n � 25; P � 0.01).(F) The decay kinetics of the NMDAR portion of the current induced by2P-uncaging of MNI-GLU (as determined in E) is plotted against the 2P-AMPA�NMDA ratio obtained from the same spines (elicited following the sameprocedure as in Fig. 3).

    Béı̈que et al. PNAS � December 19, 2006 � vol. 103 � no. 51 � 19539

    NEU

    ROSC

    IEN

    CE

    Dow

    nloa

    ded

    by g

    uest

    on

    June

    8, 2

    021

  • least with respect to synaptic AMPAR insertion) normally in theabsence of PSD-95. These results therefore suggest an intriguingselectivity in the actions of PSD-95 toward a specific population ofsynapses. Whether this apparent specificity reflects a functioninherent to PSD-95 or rather a compensatory mechanisms afterPSD-95 genetic deletion is unclear. Other members of theMAGUK superfamily of proteins might be compensating for thelack of PSD-95 in a subset of synapses.

    The decay kinetics of NMDAR currents were slower in KO mice.This effect can be accounted for, at least in part, by a greatersynaptic expression of NR2B-containing NMDAR. In particular,our results suggest that the normal, developmentally regulated,switch of primarily NR2B- to NR2A-containing synaptic NMDARis hindered in KO mice. Interestingly, the subunit composition ofNMDARs has been suggested to decisively impact the polarity ofsynaptic plasticity, although contradictory findings are found in theliterature (26–30). Because LTP is greater, and NMDARs con-taining the NR2B subunit appear to dominate in the KO mice, ourresults are broadly consistent with the idea that NR2B-containingNMDARs preferentially trigger LTP (27). However, we did notfind any evidence that the silent synapses found in KO mice wereenriched in NR2B-containing NMDARs. Because LTP is believedto be preferentially induced on silent synapses (15, 16, 23), thesefindings do not support the idea that NR2B-containing NMDARspreferentially link to the induction of LTP in KO mice. However,it is possible that spines with higher NR2B-containing NMDAR inKO mice are the ones that are more amenable to LTP regardlessof their initial AMPAR content (i.e., regardless of whether they aresilent or not) in KO mice. Our difficulty in inducing reliable LTPby 2P-uncaging unfortunately precluded us from directly testingthese possibilities at this point.

    Because thin, filopodia-like spines express less AMPARs on theirsurface than larger, more mature dendritic spines (18), we hadinitially hypothesized that the silent synapses observed in KO micewould preferentially be located on smaller spines. However, com-bined 2P-uncaging and spine morphological analysis revealed notonly that spine volume were unchanged in KO mice but also thatthe silent synapses encountered in these mice were distributed onspines independently of their volume. These data thus suggest thatsynapse functional maturation and spine morphological develop-ment can be dissociated.

    The number of AMPARs expressed at synaptic sites is believedto be governed by both activity-dependent (mainly NMDAR-dependent) and activity-independent processes (22, 26, 31, 32).Although results from this and other studies (6, 8–10) converge inshowing that synaptic AMPARs are controlled by PSD-95, themolecular details of this effect are still unclear. First, the enhance-ment of AMPAR function by transient overexpression of PSD-95

    does not require synaptic activity (10), thereby suggesting that thisrole of PSD-95 does not merely result from a general facilitatoryaction on NMDAR-dependent signaling. The observation that LTPwas not abolished (it was actually enhanced) in KO mice supportsthis view while also arguing against a direct role of PSD-95 in therecruitment process of AMPAR to synapses during NMDAR-dependent LTP. Yet, it was also found that the transient overex-pression of PSD-95 mimicked and occluded LTP (6, 10), therebysuggesting that LTP and PSD-95 control synaptic AMPAR bycommon mechanisms. Thus, it appears that PSD-95 is sufficient,although clearly not necessary, to enhance AMPAR number atsynapses. The data presented here demonstrate that PSD-95 playsa critical role in controlling AMPAR expression at a subset ofsynapses. It is difficult to infer from our data whether the lack ofAMPAR at these synapses reflects a synaptic insertion deficit orrather a synaptic stabilization deficit. Because the actual process ofinsertion of AMPARs after an LTP protocol is not compromisedin KO mice, it is possible that PSD-95 mainly plays a role inmaintaining the stability of synaptic AMPARs.

    MethodsGeneration of PSD-95-Deficient Mouse. PSD-95 gene deletion wasobtained by standard homologous recombination technique. Formore details, see Supporting Methods.

    Electrophysiology. Whole-cell recordings or field recordingswere obtained in the CA1 region of rat or mice hippocampus(P7–P25). For more details, see Supporting Methods.

    2P Uncaging and Imaging. Whole-cell recordings of CA1 pyramidalneurons were obtained by following standard procedures, exceptthat intracellular solution was supplemented with 40 �M AlexaFluor 594 to outline neuronal morphology. Slices were bathed innormal Ringer’s solution supplemented with MNI-GLU. A sectionof a neuron was imaged, and the tip of a spine was illuminated (0.8msec) with a laser tuned at 720 nm while monitoring the electro-physiological responses. For more details, see Supporting Methods.

    We thank Elise Shumsky and Dr. Zbigniew Iwinski (Carl Zeiss Micro-Imaging group) for their continuous technical assistance, Dr. HaruoKasai (University of Tokyo, Tokyo, Japan) for helpful discussion onspine volume calculation, Jason Tien for invaluable expertise in devel-oping the spine volume calculation algorithm, Dr. Gareth Thomas forcritical reading of the manuscript, and Dr James Trimmer (University ofCalifornia, Davis, CA) for providing a series of anti-PSD-95 polyclonalantibodies. This work was supported by the Howard Hughes MedicalInstitutes (R.L.H.), by grants from the National Institutes of Health (toR.L.H.), the National Alliance for Research on Schizophrenia andDepression (to J.-C.B.), and the Mental Illness Research Association (toJ.-C.B.).

    1. Malenka RC, Nicoll RA (1999) Science 285:1870–1874.2. Kornau HC, Schenker LT, Kennedy MB, Seeburg PH (1995) Science 269:1737–1740.3. Hunt CA, Schenker LJ, Kennedy MB (1996) J Neurosci 16:1380–1388.4. Sheng M, Wyszynski M (1997) BioEssays 19:847–853.5. Kennedy MB (1997) Trends Neurosci 20:264–268.6. Stein V, House DR, Bredt DS, Nicoll RA (2003) J Neurosci 23:5503–5506.7. Schnell E, Sizemore M, Karimzadegan S, Chen L, Bredt DS, Nicoll RA (2002) Proc

    Natl Acad Sci USA 99:13902–13907.8. El-Husseini AE, Schnell E, Chetkovich DM, Nicoll RA, Bredt DS (2000) Science

    290:1364–1368.9. Beique JC, Andrade R (2003) J Physiol 546:859–867.

    10. Ehrlich I, Malinow R (2004) J Neurosci 24:916–927.11. Migaud M, Charlesworth P, Dempster M, Webster LC, Watabe AM, Makhinson M,

    He Y, Ramsay MF, Morris RG, Morrison JH, et al. (1998) Nature 396:433–439.12. Hsia AY, Malenka RC, Nicoll RA (1998) J Neurophysiol 79:2013–2024.13. Ungless MA, Whistler JL, Malenka RC, Bonci A (2001) Nature 411:583–587.14. Sans N, Petralia RS, Wang YX, Blahos J, II, Hell JW, Wenthold RJ (2000) J Neurosci

    20:1260–1271.15. Liao D, Hessler NA, Malinow R (1995) Nature 375:400–404.16. Isaac JT, Nicoll RA, Malenka RC (1995) Neuron 15:427–434.17. Raastad M, Storm JF, Andersen P (1992) Eur J Neurosci 4:113–117.18. Matsuzaki M, Ellis-Davies GC, Nemoto T, Miyashita Y, Iino M, Kasai H (2001) Nat

    Neurosci 4:1086–1092.

    19. Vicini S, Wang JF, Li JH, Zhu WJ, Wang YH, Luo JH, Wolfe BB, Grayson DR (1998)J Neurophysiol 79:555–566.

    20. Flint AC, Maisch US, Weishaupt JH, Kriegstein AR, Monyer H (1997) J Neurosci17:2469–2476.

    21. Williams K, Russell SL, Shen YM, Molinoff PB (1993) Neuron 10:267–278.22. Zhu JJ, Malinow R (2002) Nat Neurosci 5:513–514.23. Matsuzaki M, Honkura N, Ellis-Davies GC, Kasai H (2004) Nature 429:761–766.24. Kopec CD, Li B, Wei W, Boehm J, Malinow R (2006) J Neurosci 26:2000–2009.25. Bagal AA, Kao JP, Tang CM, Thompson SM (2005) Proc Natl Acad Sci USA

    102:14434–14439.26. Barria A, Malinow R (2005) Neuron 48:289–301.27. Liu L, Wong TP, Pozza MF, Lingenhoehl K, Wang Y, Sheng M, Auberson YP, Wang

    YT (2004) Science 304:1021–1024.28. Massey PV, Johnson BE, Moult PR, Auberson YP, Brown MW, Molnar E,

    Collingridge GL, Bashir ZI (2004) J Neurosci 24:7821–7828.29. Weitlauf C, Honse Y, Auberson YP, Mishina M, Lovinger DM, Winder DG (2005)

    J Neurosci 25:8386–8390.30. Berberich S, Punnakkal P, Jensen V, Pawlak V, Seeburg PH, Hvalby O, Kohr G

    (2005) J Neurosci 25:6907–6910.31. Turrigiano GG, Nelson SB (2004) Nat Rev Neurosci 5:97–107.32. Luthi A, Schwyzer L, Mateos JM, Gahwiler BH, McKinney RA (2001) Nat Neurosci

    4:1102–1107.

    19540 � www.pnas.org�cgi�doi�10.1073�pnas.0608492103 Béı̈que et al.

    Dow

    nloa

    ded

    by g

    uest

    on

    June

    8, 2

    021