<<

View metadata, citation and similar papers at core.ac.uk brought to you by CORE

provided by Elsevier - Publisher Connector Perspective

Homeostatic Signaling and the Stabilization of Neural Function

Graeme W. Davis1,* 1Department of Biochemistry and Biophysics, University of California, San Francisco, San Francisco, CA 94158, USA *Correspondence: [email protected] http://dx.doi.org/10.1016/j.neuron.2013.09.044

The brain is astonishing in its complexity and capacity for change. This has fascinated scientists for more than a century, filling the pages of this journal for the past 25 years. But a paradigm shift is underway. It seems likely that the plasticity that drives our ability to learn and remember can only be meaningful in the context of otherwise stable, reproducible, and predictable baseline neural function. Without the existence of potent mechanisms that stabilize neural function, our capacity to learn and remember would be lost in the chaos of daily experiential change. This underscores two great mysteries in neuroscience. How are the functional properties of individual and neural circuits stably maintained throughout life? And, in the face of potent stabilizing mechanisms, how can neural circuitry be modified during neural development, learning, and memory? Answers are emerging in the rapidly developing field of homeostatic plasticity.

Introduction intrinsic firing properties that were characteristic of that cell It has become clear that homeostatic signaling systems act in vivo. The effect was shown to be both activity and throughout the central and peripheral nervous systems to stabi- dependent. This phenomenon has now been extended to the lize the active properties of nerve and muscle (Davis, 2006; function of entire neural circuits in both invertebrates (Haedo Marder, 2011; Turrigiano, 2011). Evidence for this has accumu- and Golowasch, 2006; see Figure 2) and the developing verte- lated by measuring how nerve and muscle respond to the persis- brate spinal cord (Gonzalez-Islas and Wenner, 2006). tent disruption of synaptic transmission, ion channel function, or A related phenomenon has been revealed in animals harboring neuronal firing. In systems ranging from Drosophila to human, mutations in genes that encode ion channels. Several studies cells have been shown to restore baseline function in the provide evidence that deleting an ion channel gene invokes continued presence of these perturbations by rebalancing ion compensatory changes in the expression of other ion channels channel expression, modifying receptor traf- in both invertebrate and mammalian systems (MacLean et al., ficking, and modulating neurotransmitter release (Frank, 2013; 2003; Swensen and Bean, 2005; Muraro et al., 2008; Andra´ sfalvy Maffei and Fontanini, 2009; Watt and Desai, 2010). In each et al., 2008; Nerbonne et al., 2008; Van Wart and Matthews, example, if baseline function is restored in the continued pres- 2006; Bergquist et al., 2010). In many cases, ion channel expres- ence of a perturbation, then the underlying signaling systems sion is rebalanced and cell-type-specific firing properties are are considered homeostatic (Figure 1). restored (Figure 2). One conclusion is that homeostatic signaling This is a rapidly growing field of investigation that can be sub- systems enable a neuron to compensate for the absence of an divided into three areas that are defined by the way in which a ionic current and re-establish cell-type-specific firing properties cell responds to activity perturbation, including the homeostatic through altered expression of other ion channels. A second control of intrinsic excitability, neurotransmitter receptor expres- conclusion is that ion channel expression is not a fixed param- sion, and presynaptic neurotransmitter release. Each area is eter associated with cell fate. Rather, a given cell type can main- introduced below. An exciting prospect is that the logic of tain characteristic firing properties using different combinations homeostatic signaling systems, if not specific molecular path- of ion channel densities. The homeostatic rebalancing of ways, will be evolutionarily conserved. The nervous systems of ion channel expression is astonishing, in part, because of its all organisms confront perturbations ranging from genetic and staggering complexity (Marder and Prinz, 2002). There can be developmental errors to changing environmental conditions. In thousands of synaptic inputs and dozens of different channels this relatively short Perspective, it is not possible to achieve a controlling the firing properties of an individual cell. The molecu- comprehensive description of the molecular advances in each lar mechanisms that achieve the homeostatic rebalancing of ion system. Rather, an attempt is made to draw parallels across channel expression remain virtually unknown (but see Muraro systems where conserved processes are emerging. et al., 2008; Driscoll et al., 2013; Temporal et al., 2012; Khorkova and Golowasch, 2007). The Homeostatic Control of Intrinsic Excitability The homeostatic control of intrinsic excitability was brought to Homeostatic Control of Synaptic Efficacy: Synaptic the forefront by experiments that followed the fate of a neuron Scaling at the Postsynaptic Density that was removed from its circuit and placed in isolated cell cul- Synaptic scaling was revealed by experiments examining the ture (Turrigiano et al., 1994). Over a period of days, the isolated effects of chronic activity suppression in cultured mammalian neuron rebalanced ion channel surface expression and restored neurons (Turrigiano et al., 1998; O’Brien et al., 1998). It is now

718 Neuron 80, October 30, 2013 ª2013 Elsevier Inc. Neuron Perspective

2008) can be sufficient to induce synaptic scaling (Figure 3). Even more remarkable is evidence that synaptic scaling can be input specific (Deeg and Aizenman, 2011) and even specific (Be´ ı¨que et al., 2011) when the manipulation of neural activity is restricted to subsets of inputs contacting a given post- synaptic neuron (Figure 3). It is not yet clear whether the magni- tude of the scaling response is matched to the magnitude of the perturbation. This is impossible to determine in experiments using tetrodotoxin (TTX) to block neural activity. In experiments in which neural activity is modulated, synaptic scaling partici- pates in the restoration of baseline firing properties in vivo (Keck et al., 2013; Hengen et al., 2013). However, synaptic scaling is often observed to act in concert with other compensa- tory changes including changes in presynaptic neurotransmitter release (Burrone et al., 2002; Kim and Tsien, 2008; Lu et al., 2013) or intrinsic excitability (Lambo and Turrigiano, 2013). It remains entirely unknown how multiple homeostatic effectors are coordi- nated to restore cell-type-specific firing properties.

Homeostatic Control of Presynaptic Neurotransmitter Release Figure 1. Evidence for the Homeostatic Control of Cellular Excitation The homeostatic modulation of presynaptic neurotransmitter Top: the firing properties of central neurons are determined by a balance of release was brought to the forefront through studies at the synaptic excitation (red vesicles and red receptors), synaptic inhibition (blue genetically tractable Drosophila neuromuscular junction (NMJ; vesicles and blue receptors), and the densities of ion channels that either mediate cellular depolarization (red ovals) or oppose cellular depolarization Davis and Goodman, 1998a). Genetic manipulations that alter (blue ovals). In response to chronic suppression of neural activity, central postsynaptic glutamate receptor function (Petersen et al., neurons can alter the relative abundance of ion channels and receptors at the 1997; Davis et al., 1997; Frank et al., 2006), muscle innervation cell surface to re-establish a set point level of activity. Bottom: at the neuro- (Davis and Goodman, 1998b), or muscle excitability (Paradis muscular junction (NMJ), chronic impairment of postsynaptic neurotransmitter receptor sensitivity or receptor abundance leads to a compensatory increase et al., 2001) were shown to induce large compensatory changes in presynaptic neurotransmitter release that precisely counteracts the change in presynaptic neurotransmitter release that precisely restore set in receptor function to achieve normal synaptic depolarization of the muscle. point muscle depolarization in response to nerve stimulation. Modified from Davis (2006). This phenomenon has been referred to as ‘‘synaptic homeosta- sis’’ but will be referred to here as ‘‘presynaptic homeostasis.’’ clear, both in vitro and in vivo, that chronic manipulation of neural This form of homeostatic signaling is evolutionarily conserved activity drives counteracting changes in neurotransmitter recep- from fly to human at the NMJ (Cull-Candy et al., 1980; Plomp tor abundance that help to restore neural activity to baseline et al., 1992). As with other forms of homeostatic plasticity, levels (Thiagarajan et al., 2005; Zhao et al., 2011; Garcia-Bere- this is a quantitatively accurate form of neuromodulation guiain et al., 2013; Mrsic-Flogel et al., 2007; Echegoyen et al., (Figure 2B; Frank et al., 2006). It can be induced in seconds to 2007; Deeg and Aizenman, 2011; Gainey et al., 2009; Keck minutes, during which its expression is independent of transcrip- et al., 2013; Hengen et al., 2013; see also Tyagarajan and Frit- tion or translation (Frank et al., 2006). It can also be stably schy, 2010). The bidirectional modulation of neurotransmitter maintained, a process that requires transcription (Marie et al., receptor abundance was initially termed ‘‘synaptic scaling’’ 2010). Presynaptic homeostasis at the NMJ is bidirectional and because the measured amplitudes of spontaneous miniature can be synapse specific (Davis and Goodman, 1998b; Daniels release events are modified in a multiplicative manner, presum- et al., 2006). ably through proportional changes in receptor abundance at Importantly, presynaptic homeostasis has also been observed every individual synapse (Turrigiano et al., 1998; Turrigiano, at mammalian central in vitro in response to differ- 2011; see also Kim and Tsien, 2008). This effect has the property ences in target innervation (Liu and Tsien, 1995), altered post- of preserving the relative differences in efficacy among the synaptic excitability (Burrone et al., 2002; Thiagarajan et al., numerous synapses on a single postsynaptic target. Because 2005; but see Goold and Nicoll, 2010; Deeg and Aizenman, of this, it has been proposed that synaptic scaling stabilizes 2011), and after chronic inhibition of neural activity (Kim and neuronal excitability while preserving learning-related informa- Ryan, 2010; Zhao et al., 2011). Regardless of the system being tion contained in relative synaptic weights. Synaptic scaling studied, the expression of presynaptic homeostasis is remark- can be induced over the course of hours to days (Turrigiano, able because it involves the rapid, persistent, and accurate mod- 2011). It requires transcription and in some cases may be ulation of presynaptic vesicle fusion. achieved through local protein translation (Sutton et al., 2006, 2007). Synaptic scaling is generally studied in response to alter- Homeostatic Design ations in global neural activity. However, manipulating the activ- The homeostatic modulation of neural function is distinct from ity of individual neurons (Goold and Nicoll, 2010; Ibata et al., other forms of neural plasticity because it is a quantitatively

Neuron 80, October 30, 2013 ª2013 Elsevier Inc. 719 Neuron Perspective

accurate form of modulation. For example, the homeostatic re- balancing of ion channel expression precisely counteracts the loss of the Kv4.2 potassium channel in pyramidal neurons and achieves firing properties that are almost indistinguishable from controls (Figure 2A). It should be pointed out, however, that compensation is not perfect because it is constrained by the unique subcellular localization and functional properties of the compensating ion channels (see also Bergquist et al., 2010). In Kv4.2 knockout pyramidal neurons, somatic excitability is precisely restored but remain hyperexcited (Chen et al., 2006; Nerbonne et al., 2008; see also Van Wart and Mat- thews, 2006). Another example of quantitative accuracy is found at the NMJ. The magnitude of postsynaptic glutamate receptor inhibition is accurately offset, over a wide range, by a graded increase in pre- synaptic neurotransmitter release (Figure 2B). The accurate modulation of presynaptic release is apparent when measured over a 10-fold range of extracellular calcium (0.3–3 mM; Figure 2C). A similarly accurate modulation of presynaptic release is observed following muscle-specific expression of an inward rectifying potassium channel (Kir2.1), which induces a nonlinear disruption of excitability because Kir2.1 inactivates during excitatory postsynaptic potential (EPSP) depolarization. Nonetheless, a precise increase in presynaptic release offset the disruption of muscle excitation caused by Kir2.1 expression and restored peak EPSP amplitude to control levels (Paradis et al., 2001). Again, compensation is accurate but imperfect since EPSP decay remains more rapid than controls, which will alter summation during a stimulus train (Paradis et al., 2001), an effect similar to that observed at the NMJ of lobster (Pulver Figure 2. Accurate Restoration of Cellular Activity through Homeostatic Signaling et al., 2005). Other examples of accurate compensation are high- (A) Sample traces from cortical pyramidal neurons from wild-type and Kv4.2 lighted in Figures 2D and 2E. One of the greatest challenges in knockout mice (Nerbonne et al., 2008, data from Figure 8 therein). The the field of homeostatic signaling is to define how accurate mod- knockout mice lack the Kv4.2 protein and current. Although acute pharma- cological inhibition severely potentiates neuronal excitability, homeostatic ulation achieved. rebalancing of potassium channel expression accurately restores firing There are several features that are commonly employed in properties to wild-type levels. both natural and engineered homeostatic signaling systems (B) Data are shown for recordings made at the Drosophila NMJ. Presynaptic that achieve quantitative accuracy (Stelling et al., 2004). First, release (quantal content) is plotted against spontaneous miniature amplitudes (mEPSP). Each data point is the average data from a single NMJ from control homeostatic systems require a set point that precisely defines NMJ (open black) or NMJ to which philanthotoxin 433 (PhTX) was applied for the output of the system. This is also the state to which the sys- 10 min prior to recording (open red). The line represents ideal homeostatic tem returns after a perturbation. Second, homeostatic systems compensation where any change in mEPSP is offset by an identical percent change in quantal content. The modulation of presynaptic release accurately generally require feedback control to precisely retarget the sys- offsets a broad range of postsynaptic perturbation. tem set point. Homeostatic systems require sensors that detect (C) Data are presented for the Drosophila NMJ plotting excitatory postsynaptic a given perturbation. By analogy, with engineered systems, it is current (EPSC) amplitude versus extracellular calcium concentration. Larvae treated with PhTx (wt + PhTX) accurately retarget control (wt) EPSC ampli- hypothesized that homeostatic signaling systems will require tudes across an order of magnitude change in extracellular calcium. Animals an error signal, representing the difference between the system harboring a loss of function mutation in rim show reduced EPSCs at all calcium set point and steady-state activity reported by the sensors. concentrations. Application of PhTX to rim mutant larvae demonstrates a Finally, the error signal is used to promote a change in homeo- failure of homeostatic compensation at all calcium concentrations (Mu¨ ller et al., 2012). static effectors that drive compensatory changes in the process (D) Intracellular recordings from a stomatogastric neuron in the intact ganglion being studied. These signaling features are often invoked in (control), following removal of the ganglion and placement in organ culture for discussions of neuronal homeostatic signaling. However, at a 10 min (Decentralized) and after 4 days in culture (4 days). After 4 days, the firing properties of the identified neuron are remarkably similar to that molecular level, our understanding remains rudimentary. The observed in the intact animal. Scale bars, 1 s/10 mV. Data modified from challenge is to begin assembling an emerging molecular ‘‘parts Khorkova and Golowasch (2007). list’’ into complete homeostatic signaling system(s) that can (E) Example traces from Xenopus central neurons including control and a explain how quantitatively control of neural activity is achieved. neuron expressing transgenic Kir2.1 (Pratt and Aizenman, 2007). Expression of Kir2.1 induces a change in the underlying current densities, including the sodium current (quantified at right), that help sustain firing properties at control Establishing a Homeostatic Set Point levels. A set point is operationally defined as the physiological state that is held constant by a homeostatic signaling system. It seems that

720 Neuron 80, October 30, 2013 ª2013 Elsevier Inc. Neuron Perspective

This work proposes that individual neurons can express different combinations of ion channels and synaptic strengths in order to arrive at cell-type-specific firing properties. Given the diversity of channels and synapses that participate in neuronal firing, the number of physiologically relevant combination channel den- sities and synaptic weights that can generate a specific firing pattern is demonstrated to be vast in silico (Prinz et al., 2004; Marder, 2011). Experimental evidence for this type of variation includes the demonstration that anatomically identical cells can have similar firing properties that are driven by diverse com- binations of underlying current densities and synaptic weights (Swensen and Bean, 2005; Schulz et al., 2006, 2007; Andra´ sfalvy et al., 2008; Goaillard et al., 2009; Temporal et al., 2012). Mechanistically, it is clear that altered ion channel transcrip- tion is involved in the homeostatic rebalancing of ion channel expression (Schulz et al., 2007; Bergquist et al., 2010). Inter- esting data from the lobster stomatogastric system has shown that neuromodulators influence the transcription of ion channels Figure 3. Cell-Autonomous and Synapse-Specific Homeostatic Plasticity in a coordinated fashion (Khorkova and Golowasch, 2007; Tem- (A) Experimental configuration is diagrammed for stimulation and simulta- poral et al., 2012). These data not only highlight the importance neous recording from adjacent CA1 pyramidal neurons in hippocampal of neuromodulation but provide insight into how the homeostatic organotypic culture that are either untransfected (Record:Ctrl, black) or transfected with channel rhodopsin 2 and photostimulated in slice culture for rebalancing of ion channel expression might be constrained. 24 hr, 50 ms light pulses at 3 Hz (Record: Photostim, blue). Below, represen- Another idea is that ion channel translation could also be a tative data are shown for AMPA-mediated currents. Below is a scatter plot of key modulatory step, downstream of the terminal selector. For recording pairs with the mean shown in red. Photostimulation causes a example, a homeostatic change in sodium channel expression decrease in AMPA current due to downscaling and a decrease in synapse number (Goold and Nicoll, 2010). after chronic manipulation of synaptic activity requires the trans- (B) Two neighboring spines with or without overlay of the Syn-YFP terminals lational regulator pumillio, a mechanism that is conserved in both from a Syn-YFP/Kir2.1- overexpressing cell. 2P uncaging of MNI-glutamate flies and mice (Driscoll et al., 2013). Finally, it is also well estab- was elicited at the tip of these spines (yellow crossed lines) and the resulting AMPAR-mediated synaptic current (2P-EPSC) is shown (Vh = À60 mV) (Be´ ı¨que lished that extrinsic factors can influence cell phenotype, one et al., 2011). A postsynaptic potentiation is seen in response to synapse example being neurotransmitter switching (Dulcis et al., 2013). specific presynaptic expression of Kir2.1. Data and images taken from Goold It remains possible that ion channel rebalancing reflects a similar and Nicoll (2010) (A) and Be´ ı¨que et al. (2011) (B). phenotypic switch, albeit more complex. Ultimately, even though we are gaining information about how a cell rebalances ion chan- the establishment of a set point of neuronal activity must be nel expression, a clear model for how the genome defines a cell- related to the specification of cell identity. For example, the firing type-specific set point for neural activity remains elusive. properties of a neuron can be as diagnostic of cell identity as any other anatomical attribute including cell size, shape, or Sensors the biochemical choice of neurotransmitter. Yet, as emphasized How cells detect a change in neural activity to initiate homeostat- above, ion channel expression, which shapes intrinsic excit- ic plasticity remains unknown. Homeostatic signaling can be ability and neural activity, is not a fixed parameter associated induced cell autonomously (Goold and Nicoll, 2010; Burrone with cell fate. How then is a set point for neuronal activity et al., 2002) and through focal application of TTX to the soma determined? (Ibata et al., 2008). These data are consistent with a somatic It is well established that combinatorial transcription factor sensor of cell-wide activity. As expected, calcium-dependent codes specify cell fate in the nervous system (Jessell, 2000). signaling is essential. For example, both synaptic upscaling Data from C. elegans suggest that cell fate is subsequently main- and downscaling have been shown to require the activity of tained through the action of ‘‘terminal selector’’ transcription fac- CamKK and CamKIV (Goold and Nicoll, 2010; Ibata et al., tors (Hobert, 2011). Terminal selectors are expressed throughout 2008). But the link between altered activity and the induction of life and control the expression of effector genes that define a homeostatic response still remains unclear. Many experiments cellular identity, including ion channels. If a terminal selector is utilize dramatic activity alterations, either blocking activity with deleted, cell fate is not maintained (Hobert, 2011). Perhaps, TTX or inducing seizure-like network activity, which will invoke rather than rigidly controlling ion channel expression, a terminal changes in calcium-dependent signaling and transcription. selector defines which ion channels can be expressed but allows However, there are examples in which moderate and graded channel expression levels to vary according to homeostatic changes in neural activity and muscle depolarization are de- feedback. tected. For example, the magnitude of glutamate receptor inhibi- The idea of a terminal selector remains consistent with tion at the NMJ is precisely offset by an increase in presynaptic groundbreaking theoretical and computational work from the release (Figures 1 and 2) implying a sensor that is able to detect stomatogastric ganglion examining how a set point is retargeted graded changes in muscle excitation. Similarly, moderate through homeostatic feedback (Prinz et al., 2004; Marder, 2011). changes in neuronal firing measured in visual cortex after visual

Neuron 80, October 30, 2013 ª2013 Elsevier Inc. 721 Neuron Perspective

deprivation can invoke homeostatic plasticity, leading to the was shown to stimulate expression of Homer1a, which subse- restoration of baseline firing rates (Keck et al., 2013; Hengen quently activates mGluR signaling in an agonist-independent et al., 2013; see also Deeg and Aizenman, 2011). Early efforts manner (Hu et al., 2010). This model is intriguing because the to model homeostatic plasticity in the stomatogastric system control of mGluR subcellular localization has the potential to have emphasized that multiple activity sensors are necessary define the spatial extent of the homeostatic response. In a sepa- to discriminate quantitative differences in neuronal firing (Liu rate set of studies, enhanced network activity induces Plk2, et al., 1998). Yet, biologically, a system of coordinated sensors which phosphorylates the postsynaptic scaffolding protein with the fidelity to follow neural activity remains unknown. SPAR in a CDK5-dependent manner. Subsequent SPAR degra- An interesting possibility is that metabolic sensors may be em- dation weakens the retention of AMPA receptors at the postsyn- ployed in addition to, or in parallel with, changes in intracellular aptic membrane, facilitating synaptic downscaling (Seeburg calcium. In dissociated hippocampal culture, eukaryotic elonga- et al., 2008; Seeburg and Sheng, 2008). Finally, although not tion factor 2 (eEF2) has been implicated as a sensor that can an immediate early gene, retinoic acid has been shown to be detect disruption of glutamatergic transmission (Sutton et al., required for synaptic upscaling, in this case following postsyn- 2004, 2007). Additional work implicates a function for TOR- aptic glutamate receptor inhibition (Wang et al., 2011; Sarti dependent signaling downstream of AMPA receptor blockade et al., 2012). In this model a decrease in dendritic calcium after (Henry et al., 2012). The potential importance of this signaling AMPA receptor blockade induces retinoic acid synthesis and system for homeostasis in vivo is emphasized in experiments subsequent AMPA receptor production (Wang et al., 2011). Ret- demonstrating that TOR signaling is essential for balanced inoic acid acts via the retinoic acid receptor (RAR-a)(Sarti et al., network excitation and inhibition (Bateup et al., 2013). The 2012) and could, potentially, signal cell autonomously (Wang importance of TOR is also emphasized by work at the Drosophila et al., 2011). NMJ in vivo, showing that genetic disruption of TOR and S6 Other advances center on how surface delivery and synaptic Kinase signaling blocks the sustained expression of presynaptic retention of AMPA receptors is controlled so that a homeostatic homeostasis (Penney et al., 2012). In many systems, TOR response can be graded and potentially matched to the magni- signaling is used to detect qualitative changes in the cellular tude of a perturbation. For example, PICK1 (protein interacting environment and, thereby, regulates cellular homeostasis and with C-kinase) scaffolds an intracellular AMPA receptor pool. growth (Laplante and Sabatini, 2012). As such, it is a candidate There is evidence that PICK1 levels are decreased in a graded for detecting quantitative changes in neural function and stimu- fashion in response to chronic activity inhibition, releasing lating downstream homeostatic plasticity. AMPA receptors for translocation to the plasma membrane (Anggono et al., 2011). Other work focuses on how AMPA recep- Homeostatic Effectors: Synaptic Scaling tors are retained at the postsynaptic density by PSD95, PSD93, Synaptic scaling is expressed as a change in neurotransmitter and SAP102. It has been shown that PSD95 and SAP102 levels receptor abundance. Although a great deal has been discovered are modulated bidirectionally by neural activity (Sun and Turri- about the transcription, assembly, and trafficking of glutamate giano, 2011). In this study, PSD95 is shown to be necessary receptors, the mechanisms that control receptor trafficking in a but not sufficient for synaptic scaling, acting through the regu- homeostatic fashion remain largely unknown. Many key issues lated organization of the postsynaptic scaffold (Sun and Turri- remain to be molecularly defined, including how synaptic scaling giano, 2011). Clearly, there will be additional complexity as an is achieved in a cell-wide fashion with proportional effects at increasing number of molecules are shown to be necessary for every active zone. Similarly, there is very little information to synaptic scaling including MHC1 (Goddard et al., 2007), BDNF explain how the synaptic scaling mechanism becomes limited (Rutherford et al., 1998; Jakawich et al., 2010; Correˆ a et al., as neuronal firing properties are restored toward baseline, set 2012), and Beta3-integrins (McGeachie et al., 2011). point levels, and how the system is eventually shut off (but see Tatavarty et al., 2013). In attempting to define how the homeo- Homeostatic Effectors: Presynaptic Homeostasis static control of glutamate receptor trafficking is achieved, it is The power of model system forward genetics in Drosophila has useful to make comparisons to nonneuronal systems in which opened the door to a mechanistic understanding of presynaptic homeostatic control of surface transporters and ion channels homeostasis. An -based forward genetic has been defined without the added complexity of cell diversity. screen is ongoing, based on intracellular recordings of neuro- These systems include trafficking of ENaC channels during the muscular transmission, to identify mutations that prevent the ho- systemic control of salt balance and trafficking of the Glut4 meostatic enhancement of presynaptic neurotransmitter release glucose transporter during glucose homeostasis (Lifton et al., after pharmacological inhibition of postsynaptic glutamate 2001; Martin et al., 2006; Leto and Saltiel, 2012). Several ad- receptors (Dickman and Davis, 2009; Mu¨ ller et al., 2011; Younger vances are highlighted here that provide insight into emerging et al., 2013). To date, more than 1,000 mutations and RNAi have homeostatic control of glutamate receptor trafficking. been tested (Dickman and Davis, 2009; Mu¨ ller et al., 2011; The induction of synaptic scaling has been an area of consid- Younger et al., 2013). Based largely on the results of this forward erable progress. An emerging theme is the activity-dependent genetic approach, a model has emerged to explain how synaptic induction of immediate early gene signaling including Homer1a, vesicle release is precisely potentiated at the NMJ. Arc (Arg3.1), Narp, and Polo-like kinase 2 (Plk2) (Seeburg et al., Two presynaptic processes converge to potentiate vesicle 2008; Hu et al., 2010; Chang et al., 2010; Be´ ı¨que et al., 2011; fusion during presynaptic homeostasis: (1) potentiation of pre- Shepherd et al., 2006). In one study, enhanced network activity synaptic calcium influx and (2) potentiation of the readily

722 Neuron 80, October 30, 2013 ª2013 Elsevier Inc. Neuron Perspective

blocks presynaptic homeostasis (Figure 2C), was particularly informative. When the homeostatic modulation of presynaptic calcium influx and RRP size were tested in RIM mutants, only the modulation of the RRP was blocked and synaptic homeosta- sis was prevented (Mu¨ ller et al., 2012). The means by which RIM mediates this activity has yet to be determined. The RIM- interacting molecules Rab3 and Rab3-GAP also participate in presynaptic homeostasis (Mu¨ ller et al., 2011). In mammalian sys- tems, these molecules establish a biochemical bridge between the calcium channel and the (Han et al., 2011; Kaeser et al., 2011). This may represent a central, regulated scaf- fold that coordinates the homeostatic modulation of the RRP with calcium entry. Additional genes have been found to be essential for presynaptic homeostasis including postsynaptic scaffolding (Pilgram et al., 2011), postsynaptic TOR/S6K (Pen- ney et al., 2012), and micro-RNA signaling (Tsurudome et al., 2010), all nicely summarized in a recent review of homeostatic plasticity at the Drosophila NMJ (Frank, 2013). Parallels have emerged at mammalian central synapses, Figure 4. Presynaptic Homeostasis Is Achieved by Parallel Changes consistent with the homeostatic modulation of both vesicle pools in the Size of the Readily Releasable Vesicle Pool and Presynaptic and presynaptic calcium influx. Chronic activity blockade has Calcium Influx (A and B) Variance-mean analysis supports a RIM-dependent modulation of been shown to cause a correlated increase in both presynaptic the RRP during synaptic homeostasis. (A) Example EPSC traces for a WT NMJ release and calcium influx, imaged simultaneously through at the indicated extracellular calcium concentration (millimolar). (B) Example coexpression of transgenic reporters for vesicle fusion and EPSC amplitude variance-mean plots of two WT synapses in the absence calcium (Zhao et al., 2011). Mechanistically, presynaptic CDK5 (gray) and presence (black) of PhTX with parabola fits (solid lines) that were extrapolated to the x intercept (dashed lines; see Mu¨ ller et al., 2012). has been implicated. Loss or inhibition of CDK5 potentiates pre- (C) Representative traces for measurement of the spatially averaged calcium synaptic release by promoting calcium influx and enhanced signal within synaptic boutons at the NMJ in a wild-type (control) and GluRIIA access to a recycling pool of synaptic vesicles. Chronic activity mutant. The homeostatic enhancement of presynaptic release is correlated with a statistically significant increase in the peak amplitude of the spatially suppression phenocopies these effects and causes a decrease averaged calcium signal (see Mu¨ ller and Davis, 2012). in synaptic CDK5 implying a causal link (Kim and Ryan, 2010). The activity of CDK5 has been shown to be balanced by calci- releasable pool (RRP) of synaptic vesicles (Figure 4). First, a neurin A and, together, these molecules act via the CaV2.2 combination of calcium imaging and genetic data demonstrate calcium channel (Kim and Ryan, 2013). Remarkably, the CDK5/ that an increase in presynaptic calcium influx through the Calcineurin-dependent modulation of presynaptic release has CaV2.1 calcium channel is necessary to achieve a homeostatic sufficient signaling capacity to cause the silencing and unsilenc- increase in vesicle release (Mu¨ ller et al., 2011, 2012). A surprising ing of individual active zones in hippocampal cultures (Kim and mechanism is employed to modulate presynaptic calcium influx. Ryan, 2013). The involvement of a presynaptic DEG/ENaC sodium leak chan- nel was uncovered in the aforementioned genetic screen. In the Enabling a Homeostatic Response through Permissive emerging model, presynaptic DEG/ENaC channel insertion at or Signaling near the nerve terminal causes low-voltage modulation of the Studies at the Drosophila NMJ and mammalian central synapses presynaptic resting potential due to sodium leak and subsequent demonstrate that secreted factors create an environment that is potentiation of presynaptic calcium influx (Figure 5). This model necessary for the expression and/or maintenance of homeostat- is attractive because it provides an analog mechanism that could ic plasticity including both presynaptic homeostasis and post- fine-tune presynaptic calcium influx according to the demands synaptic scaling. Since these factors do not dictate the timing of the homeostatic signaling system. Low-voltage modulation or magnitude of the homeostatic response, they are consid- of neurotransmitter release has been observed in systems ered essential, permissive cues. At the Drosophila NMJ, bone ranging from the crayfish NMJ to the rodent hippocampus (Woj- morphogenetic protein (BMP) signaling from muscle to moto- towicz and Atwood, 1983; Awatramani et al., 2005; Christie et al., neuron drives NMJ growth during larval development (McCabe 2011), although links to homeostatic plasticity have not been et al., 2003). Subsequently, it was demonstrated that genetic made in these systems. Interestingly, ENaC channels can be deletion of the BMP ligand, a presynaptic BMP receptor, or considered as homeostatic effector proteins during the systemic downstream transcription all blocks synaptic homeostasis control of salt balance (Lifton et al., 2001). (Goold and Davis, 2007). Importantly, BMP signaling does not Remarkably, the potentiation of presynaptic calcium influx function at the NMJ to instruct a change in neurotransmitter alone is not sufficient to drive a homeostatic change in synaptic release. Instead, BMP-dependent transcription permits the in- vesicle fusion. A parallel increase in the RRP of synaptic vesicles duction of synaptic homeostasis, which is expressed locally at is required (Weyhersmu¨ ller et al., 2011; Mu¨ ller et al., 2012). An the NMJ. The nature of the permissive signal downstream of analysis of mutations in RIM (Rab3 Interacting Molecule), which BMP-dependent transcription remains unknown (Goold and

Neuron 80, October 30, 2013 ª2013 Elsevier Inc. 723 Neuron Perspective

Figure 5. Emerging Model for the Expression of Presynaptic Homeostasis Schematic of an active zone at the Drosophila NMJ shows presynaptic CaV2.1 calcium channels (blue), the triggered calcium microdomain (red), postsynaptic glutamate re- ceptors (GluRIIA, dark gray), and presynaptic ENaC channel (pink). Two genes, pickpocket11 and pickpocket16, were discovered to be neces- sary, presynaptically, for the rapid induction and sustained expression of presynaptic homeostasis (Younger et al., 2013). These genes encode sub- units of an Epithelial/Degenerin (ENaC) sodium leak channel. These genes are cotranscribed and transcriptionally upregulated after persistent disruption of postsynaptic glutamate receptor function. These and other data support a model in which ENaC channel insertion drives a modest depolarization of the presynaptic resting potential (DV), which enhances presynaptic calcium and neurotransmitter release. A parallel change in the readily releasable pool of vesicles is also necessary for synaptic homeostasis, which relies on the presynaptic adaptor proteins RIM (Mu¨ ller et al., 2012) and RIM binding protein (RBP) (green). When both processes are enabled, a homeostatic enhancement of release is observed. A number of additional synaptic proteins have been shown to be required for presynaptic homeostasis (dark blue text). Among them, there is evidence for the involvement of micro-RNA-based signaling (Mir10, gray; Tsurudome et al., 2010) and permissive BMP-mediated signaling released from muscle (gray) and acting at the motoneuron soma (gray; Goold and Davis, 2007). Postsynaptically, there is evidence for the required function of Tor and S6K as well as Dystrobrevin-dependent scaffolding (Penney et al., 2012; Pilgram et al., 2011). Figure modified from Younger et al. (2013). Additional molecular mechanisms involved in presynaptic homeostasis have been recently reviewed (Frank, 2013).

Davis, 2007). A related function has been proposed for TNF-a at control to ensure a robust, reproducible outcome (Baumgardt mammalian central synapses. TNF-a is required for synaptic et al., 2007). From this perspective, homeostatic plasticity might scaling, is released from glia in response to prolonged activity serve to correct developmental inaccuracies but would not be blockade, and is sufficient to drive synaptic scaling (Beattie invoked as part of normal developmental signaling. Data from et al., 2002; Stellwagen et al., 2005; Stellwagen and Malenka, the Drosophila NMJ support this idea. 2006). But it was recently demonstrated that the rapid induction The NMJ of Drosophila larvae grow 100-fold in volume of synaptic scaling after 4–6 hr of activity blockade is normal in during a 5-day period of postembryonic development. In the absence of TNF-a. Synaptic scaling is only blocked if TNF- Drosophila, as in most systems, when a muscle fiber grows a is removed >12 hr prior to activity blockade. It is argued, based the input resistance drops precipitously, requiring enhanced on these data, that TNF-a is a permissive signal that maintains presynaptic release to achieve constant muscle depolarization synapses in a state amenable to homeostatic modulation (Stein- (Davis and Goodman, 1998a). However, presynaptic homeosta- metz and Turrigiano, 2010). While the relevance of permissive sis does not appear to be involved. Forward genetic screening homeostatic signaling remains to be determined, permissive has identified several mutations that block presynaptic homeo- signaling could be used to control whether or not homeostatic stasis without altering anatomical or functional neuromuscular plasticity is expressed at different times during development development (Dickman and Davis, 2009; Mu¨ ller et al., 2011; (Maffei and Fontanini, 2009; Echegoyen et al., 2007) and its in- Younger et al., 2013). This observation was extended by recent duction in the context of stress or disease (Goold and Davis, work using the ENaC channel blocker benzamil to acutely disrupt 2007; Steinmetz and Turrigiano, 2010). presynaptic homeostasis. Benzamil was applied to the NMJ of animals lacking the muscle-specific GluRIIA glutamate receptor Interface of Development and Homeostatic Plasticity subunit, a perturbation that is persistent throughout develop- The nervous system is remarkably plastic during development. ment and induces presynaptic homeostasis. Benzamil erased Individual neurons and muscle change dramatically in size and the expression of presynaptic homeostasis, leaving behind a complexity. Synaptic connectivity is refined through mecha- synapse with unpotentiated wild-type release and wild-type nisms of synaptic competition. New cells are integrated into fully anatomy (Younger et al., 2013). Together, these data demon- functioning neural circuitry. Do these changes represent pertur- strate that presynaptic homeostasis is uncoupled from the bations that are stabilized by homeostatic signaling? One mechanisms that achieve anatomical and physiological NMJ approach has been to define the cellular parameters that are growth. One possibility is that presynaptic homeostasis is only held constant during periods of developmental growth, but this induced developmentally when a cellular set point differs from has not defined whether constancy is achieved through homeo- ongoing activity. If the set point is developmentally programmed static control (Bucher et al., 2005). Answering this question is to change along with the maturation of cell fate, then a develop- likely to be important for understanding how homeostatic plas- mental change in cellular function could occur without the induc- ticity might participate in diseases including autism spectrum tion of homeostatic plasticity. disorders (ASDs) and schizophrenia that can be traced back to In the mammalian CNS, homeostatic and developmental plas- alterations in early brain development (Ramocki and Zoghbi, ticity coexist. This is nicely documented in a binocular region of 2008; Bourgeron, 2009). The construction of an embryo from a visual cortex after monocular deprivation (Mrsic-Flogel et al., single cell is a tightly choreographed process that includes 2007). When cells receive input predominantly from an open inductive signaling with both negative and positive feedback eye, deprived eye input is diminished, consistent with classical

724 Neuron 80, October 30, 2013 ª2013 Elsevier Inc. Neuron Perspective

synaptic competition. However, when cells receive input REFERENCES predominantly from the deprived eye, these inputs are strength- ened, consistent with homeostatic plasticity. Binocular depriva- Andra´ sfalvy, B.K., Makara, J.K., Johnston, D., and Magee, J.C. (2008). Altered synaptic and non-synaptic properties of CA1 pyramidal neurons in Kv4.2 tion also induces homeostatic synaptic strengthening. Although knockout mice. J. Physiol. 586, 3881–3892. these processes coexist, it remains unclear whether homeostat- Anggono, V., Clem, R.L., and Huganir, R.L. (2011). PICK1 loss of function ic plasticity normally participates in ocular dominance indepen- occludes homeostatic synaptic scaling. J. Neurosci. 31, 2188–2196. dent of an experimental perturbation such as eye suturing. In other examples, cell-autonomous suppression of neural activity Awatramani, G.B., Price, G.D., and Trussell, L.O. (2005). Modulation of trans- mitter release by presynaptic resting potential and background calcium levels. has been shown to induce changes in synaptic connectivity as Neuron 48, 109–121. well as homeostatic plasticity, but the effects are separated in Bateup, H.S., Johnson, C.A., Denefrio, C.L., Saulnier, J.L., Kornacker, K., and time (Burrone et al., 2002). Sabatini, B.L. (2013). Excitatory/inhibitory synaptic imbalance leads to hippo- campal hyperexcitability in mouse models of tuberous sclerosis. Neuron 78, Disease 510–522. There are emerging molecular links between homeostatic plas- Baumgardt, M., Miguel-Aliaga, I., Karlsson, D., Ekman, H., and Thor, S. (2007). ticity and neurological disease. The schizophrenia-associated Specification of neuronal identities by feedforward combinatorial coding. gene dysbindin was isolated in a forward genetic screen for PLoS Biol. 5, e37. mutations that block presynaptic homeostasis (Dickman and Beattie, E.C., Stellwagen, D., Morishita, W., Bresnahan, J.C., Ha, B.K., Von Davis, 2009). Homer and mGluR signaling are implicated in Zastrow, M., Beattie, M.S., and Malenka, R.C. (2002). Control of synaptic strength by glial TNFalpha. Science 295, 2282–2285. mouse models of fragile X syndrome (Ronesi et al., 2012), as is ret- inoic acid (Soden and Chen, 2010). Others have speculated the Be´ ı¨que, J.-C., Na, Y., Kuhl, D., Worley, P.F., and Huganir, R.L. (2011). Arc- involvement of disrupted homeostatic signaling in posttraumatic dependent synapse-specific homeostatic plasticity. Proc. Natl. Acad. Sci. USA 108, 816–821. epilepsy (Houweling et al., 2005), Rett syndrome (Ramocki and Zoghbi, 2008; Qiu et al., 2012), and autism spectrum disorders Bergquist, S., Dickman, D.K., and Davis, G.W. (2010). A hierarchy of cell (Bourgeron, 2009). A wealth of information is emerging regarding intrinsic and target-derived homeostatic signaling. Neuron 66, 220–234. rare de novo mutations with strong effects in autism spectrum Bourgeron, T. (2009). A synaptic trek to autism. Curr. Opin. Neurobiol. 19, disorders and it is possible that further associations with homeo- 231–234. static plasticity will emerge (Murdoch and State, 2013). Clearly, it Bucher, D., Prinz, A.A., and Marder, E. (2005). Animal-to-animal variability in will be important to advance our understanding of how homeo- motor pattern production in adults and during growth. J. Neurosci. 25, static plasticity interfaces with the function and plasticity of 1611–1619. neural circuitry in vivo, a topic that has been recently reviewed Burrone, J., O’Byrne, M., and Murthy, V.N. (2002). Multiple forms of synaptic (Watt and Desai, 2010; Maffei and Fontanini, 2009; Turrigiano, plasticity triggered by selective suppression of activity in individual neurons. Nature 420, 414–418. 2011). It will be just as important to gain a complete molecular understanding of the molecular design and implementation of Chang, M.C., Park, J.M., Pelkey, K.A., Grabenstatter, H.L., Xu, D., Linden, D.J., Sutula, T.P., McBain, C.J., and Worley, P.F. (2010). Narp regulates neuronal homeostatic signaling within individual neurons. Once homeostatic scaling of excitatory synapses on parvalbumin-expressing inter- understood, the manipulation of homeostatic signaling could neurons. Nat. Neurosci. 13, 1090–1097. enable therapeutic manipulation of neuronal activity with far Chen, X., Yuan, L.-L., Zhao, C., Birnbaum, S.G., Frick, A., Jung, W.E., reaching implications. By extension, it might become possible Schwarz, T.L., Sweatt, J.D., and Johnston, D. (2006). Deletion of Kv4.2 gene for medicine to treat the compensated nervous system, rather eliminates dendritic A-type K+ current and enhances induction of long-term than the underlying disruptions that initiate disease. potentiation in hippocampal CA1 pyramidal neurons. J. Neurosci. 26, 12143–12151.

Conclusions Christie, J.M., Chiu, D.N., and Jahr, C.E. (2011). Ca(2+)-dependent enhance- ment of release by subthreshold somatic depolarization. Nat. Neurosci. 14, This field is wide open for exploration. The excitement surround- 62–68. ing this emerging field is that it spans so many disciplines, from control theory and modeling to biophysics, developmental Correˆ a, S.A.L., Hunter, C.J., Palygin, O., Wauters, S.C., Martin, K.J., McKenzie, C., McKelvey, K., Morris, R.G., Pankratov, Y., Arthur, J.S., and biology, and human disease. How is cell-type-specific homeo- Frenguelli, B.G. (2012). MSK1 regulates homeostatic and experience-depen- static plasticity achieved? Can the phenotypic modulation of dent . J. Neurosci. 32, 13039–13051. neural activity be visualized in vivo while homeostatic plasticity Cull-Candy, S.G., Miledi, R., Trautmann, A., and Uchitel, O.D. (1980). On the is induced and executed? If so, might it be possible to identify release of transmitter at normal, myasthenia gravis and myasthenic syndrome enzymatic activities and protein interactions that precede, track, affected human end-plates. J. Physiol. 299, 621–638. or lag the phenotypic expression of a homeostatic response? Daniels, R.W., Collins, C.A., Chen, K., Gelfand, M.V., Featherstone, D.E., and These are just a few major questions, in addition to those raised DiAntonio, A. (2006). A single vesicular glutamate transporter is sufficient to fill throughout this Perspective, that remain to be experimentally a synaptic vesicle. Neuron 49, 11–16, http://dx.doi.org/10.1016/j.neuron. 2005.11.032. addressed. Davis, G.W. (2006). Homeostatic control of neural activity: from phenomenol- ACKNOWLEDGMENTS ogy to molecular design. Annu. Rev. Neurosci. 29, 307–323, http://dx.doi.org/ 10.1146/annurev.neuro.28.061604.135751.

This work is supported by NIH grant NS39313 to G.W.D. I thank Meg Younger, Davis, G.W., and Goodman, C.S. (1998a). Genetic analysis of synaptic devel- Kevin Ford, Michael Gavino, Martin Muller, and Santiago Archila for help with opment and plasticity: homeostatic regulation of synaptic efficacy. Curr. Opin. the preparation of this Perspective. Neurobiol. 8, 149–156.

Neuron 80, October 30, 2013 ª2013 Elsevier Inc. 725 Neuron Perspective

Davis, G.W., and Goodman, C.S. (1998b). Synapse-specific control of synap- Houweling, A.R., Bazhenov, M., Timofeev, I., Steriade, M., and Sejnowski, T.J. tic efficacy at the terminals of a single neuron. Nature 392, 82–86, http://dx.doi. (2005). Homeostatic synaptic plasticity can explain post-traumatic epilepto- org/10.1038/32176. genesis in chronically isolated neocortex. Cereb. Cortex 15, 834–845.

Davis, G.W., Schuster, C.M., and Goodman, C.S. (1997). Genetic analysis of Hu, J.-H., Park, J.M., Park, S., Xiao, B., Dehoff, M.H., Kim, S., Hayashi, T., the mechanisms controlling target selection: target-derived Fasciclin II regu- Schwarz, M.K., Huganir, R.L., Seeburg, P.H., et al. (2010). Homeostatic scaling lates the pattern of synapse formation. Neuron 19, 561–573. requires group I mGluR activation mediated by Homer1a. Neuron 68, 1128– 1142. Deeg, K.E., and Aizenman, C.D. (2011). Sensory modality-specific homeostat- ic plasticity in the developing optic tectum. Nat. Neurosci. 14, 548–550. Ibata, K., Sun, Q., and Turrigiano, G.G. (2008). Rapid synaptic scaling induced by changes in postsynaptic firing. Neuron 57, 819–826. Dickman, D.K., and Davis, G.W. (2009). The schizophrenia susceptibility gene dysbindin controls synaptic homeostasis. Science 326, 1127–1130, http://dx. Jakawich, S.K., Nasser, H.B., Strong, M.J., McCartney, A.J., Perez, A.S., doi.org/10.1126/science.1179685. Rakesh, N., Carruthers, C.J., and Sutton, M.A. (2010). Local presynaptic activ- ity gates homeostatic changes in presynaptic function driven by dendritic BDNF synthesis. Neuron 68, 1143–1158. Driscoll, H.E., Muraro, N.I., He, M., and Baines, R.A. (2013). Pumilio-2 regu- lates translation of Nav1.6 to mediate homeostasis of membrane excitability. Jessell, T.M. (2000). Neuronal specification in the spinal cord: inductive signals J. Neurosci. 33, 9644–9654. and transcriptional codes. Nat. Rev. Genet. 1, 20–29.

Dulcis, D., Jamshidi, P., Leutgeb, S., and Spitzer, N.C. (2013). Neurotrans- Kaeser, P.S., Deng, L., Wang, Y., Dulubova, I., Liu, X., Rizo, J., and Su¨ dhof, mitter switching in the adult brain regulates behavior. Science 340, 449–453, T.C. (2011). RIM proteins tether Ca2+ channels to presynaptic active zones http://dx.doi.org/10.1126/science.1234152. via a direct PDZ-domain interaction. Cell 144, 282–295.

Echegoyen, J., Neu, A., Graber, K.D., and Soltesz, I. (2007). Homeostatic plas- Keck, T., Keller, G.B., Jacobsen, R.I., Eysel, U.T., Bonhoeffer, T., and Hu¨ bener, ticity studied using in vivo hippocampal activity-blockade: synaptic scaling, M. (2013). Synaptic scaling and homeostatic plasticity in the mouse visual cor- intrinsic plasticity and age-dependence. PLoS ONE 2, e700. tex in vivo. Neuron 80. Published online October 16, 2013. http://dx.doi.org/ 10.1016/j.neuron.2013.08.018. Frank, C.A. (2013). Homeostatic plasticity at the Drosophila neuromuscular junction. Neuropharmacology. Published online June 24, 2013. http://dx.doi. Khorkova, O., and Golowasch, J. (2007). Neuromodulators, not activity, con- org/10.1016/j.neuropharm.2013.06.015. trol coordinated expression of ionic currents. J. Neurosci. 27, 8709–8718.

Frank, C.A., Kennedy, M.J., Goold, C.P., Marek, K.W., and Davis, G.W. (2006). Kim, S.H., and Ryan, T.A. (2010). CDK5 serves as a major control point in Mechanisms underlying the rapid induction and sustained expression of syn- neurotransmitter release. Neuron 67, 797–809. aptic homeostasis. Neuron 52, 663–677. Kim, S.H., and Ryan, T.A. (2013). Balance of calcineurin Aa and CDK5 activities Gainey, M.A., Hurvitz-Wolff, J.R., Lambo, M.E., and Turrigiano, G.G. (2009). sets release probability at nerve terminals. J. Neurosci. 33, 8937–8950. Synaptic scaling requires the GluR2 subunit of the AMPA receptor. J. Neurosci. 29, 6479–6489. Kim, J., and Tsien, R.W. (2008). Synapse-specific adaptations to inactivity in hippocampal circuits achieve homeostatic gain control while dampening Garcia-Bereguiain, M.A., Gonzalez-Islas, C., Lindsly, C., Butler, E., Hill, A.W., network reverberation. Neuron 58, 925–937. and Wenner, P. (2013). In vivo synaptic scaling is mediated by GluA2-lacking Lambo, M.E., and Turrigiano, G.G. (2013). Synaptic and intrinsic homeostatic AMPA receptors in the embryonic spinal cord. J. Neurosci. 33, 6791–6799. mechanisms cooperate to increase L2/3 pyramidal neuron excitability during a late phase of critical period plasticity. J. Neurosci. 33, 8810–8819. Goaillard, J.-M., Taylor, A.L., Schulz, D.J., and Marder, E. (2009). Functional consequences of animal-to-animal variation in circuit parameters. Nat. Neuro- Laplante, M., and Sabatini, D.M. (2012). mTOR signaling in growth control and sci. 12, 1424–1430. disease. Cell 149, 274–293.

Goddard, C.A., Butts, D.A., and Shatz, C.J. (2007). Regulation of CNS synap- Leto, D., and Saltiel, A.R. (2012). Regulation of glucose transport by insulin: ses by neuronal MHC class I. Proc. Natl. Acad. Sci. USA 104, 6828–6833. traffic control of GLUT4. Nat. Rev. Mol. Cell Biol. 13, 383–396.

Gonzalez-Islas, C., and Wenner, P. (2006). Spontaneous network activity in Lifton, R.P., Gharavi, A.G., and Geller, D.S. (2001). Molecular mechanisms of the embryonic spinal cord regulates AMPAergic and GABAergic synaptic human hypertension. Cell 104, 545–556. strength. Neuron 49, 563–575. Liu, G., and Tsien, R.W. (1995). Properties of synaptic transmission at single Goold, C.P., and Davis, G.W. (2007). The BMP ligand Gbb gates the expres- hippocampal synaptic boutons. Nature 375, 404–408. sion of synaptic homeostasis independent of synaptic growth control. Neuron 56, 109–123, http://dx.doi.org/10.1016/j.neuron.2007.08.006. Liu, Z., Golowasch, J., Marder, E., and Abbott, L.F. (1998). A model neuron with activity-dependent conductances regulated by multiple calcium sensors. Goold, C.P., and Nicoll, R.A. (2010). Single-cell optogenetic excitation drives J. Neurosci. 18, 2309–2320. homeostatic synaptic depression. Neuron 68, 512–528. Lu, W., Bushong, E.A., Shih, T.P., Ellisman, M.H., and Nicoll, R.A. (2013). The Haedo, R.J., and Golowasch, J. (2006). Ionic mechanism underlying recovery cell-autonomous role of excitatory synaptic transmission in the regulation of of rhythmic activity in adult isolated neurons. J. Neurophysiol. 96, 1860–1876. neuronal structure and function. Neuron 78, 433–439. MacLean, J.N., Zhang, Y., Johnson, B.R., and Harris-Warrick, R.M. (2003). Han, Y., Kaeser, P.S., Su¨ dhof, T.C., and Schneggenburger, R. (2011). RIM Activity-independent homeostasis in rhythmically active neurons. Neuron 37, determines Ca2+ channel density and vesicle docking at the presynaptic 109–120. active zone. Neuron 69, 304–316. Maffei, A., and Fontanini, A. (2009). Network homeostasis: a matter of coordi- Hengen, K.B., Lambo, M.E., Van Hooser, S.D., Katz, D.B., and Turrigiano, G.G. nation. Curr. Opin. Neurobiol. 19, 168–173. (2013). Firing rate homeostasis in visual cortex of freely behaving rodents. Neuron. Published online October 16, 2013. http://dx.doi.org/10.1016/j. Marder, E. (2011). Variability, compensation, and modulation in neurons and neuron.2013.08.038. circuits. Proc. Natl. Acad. Sci. USA 108(Suppl 3 ), 15542–15548.

Henry, F.E., McCartney, A.J., Neely, R., Perez, A.S., Carruthers, C.J.L., Stuen- Marder, E., and Prinz, A.A. (2002). Modeling stability in neuron and network kel, E.L., Inoki, K., and Sutton, M.A. (2012). Retrograde changes in presynaptic function: the role of activity in homeostasis. Bioessays 24, 1145–1154. function driven by dendritic mTORC1. J. Neurosci. 32, 17128–17142. Marie, B., Pym, E., Bergquist, S., and Davis, G.W. (2010). Synaptic homeosta- Hobert, O. (2011). Regulation of terminal differentiation programs in the sis is consolidated by the cell fate gene gooseberry, a Drosophila pax3/7 nervous system. Annu. Rev. Cell Dev. Biol. 27, 681–696. homolog. J. Neurosci. 30, 8071–8082.

726 Neuron 80, October 30, 2013 ª2013 Elsevier Inc. Neuron Perspective

Martin, O.J., Lee, A., and McGraw, T.E. (2006). GLUT4 distribution between Qiu, Z., Sylwestrak, E.L., Lieberman, D.N., Zhang, Y., Liu, X.-Y., and Ghosh, A. the plasma membrane and the intracellular compartments is maintained by (2012). The Rett syndrome protein MeCP2 regulates synaptic scaling. an insulin-modulated bipartite dynamic mechanism. J. Biol. Chem. 281, J. Neurosci. 32, 989–994. 484–490. Ramocki, M.B., and Zoghbi, H.Y. (2008). Failure of neuronal homeostasis McCabe, B.D., Marque´ s, G., Haghighi, A.P., Fetter, R.D., Crotty, M.L., Haerry, results in common neuropsychiatric phenotypes. Nature 455, 912–918. T.E., Goodman, C.S., and O’Connor, M.B. (2003). The BMP homolog Gbb pro- vides a retrograde signal that regulates synaptic growth at the Drosophila Ronesi, J.A., Collins, K.A., Hays, S.A., Tsai, N.-P., Guo, W., Birnbaum, S.G., neuromuscular junction. Neuron 39, 241–254. Hu, J.H., Worley, P.F., Gibson, J.R., and Huber, K.M. (2012). Disrupted Homer scaffolds mediate abnormal mGluR5 function in a mouse model of fragile X McGeachie, A.B., Cingolani, L.A., and Goda, Y. (2011). Stabilising influence: syndrome. Nat. Neurosci. 15, 431–440, S1. integrins in regulation of synaptic plasticity. Neurosci. Res. 70, 24–29. Rutherford, L.C., Nelson, S.B., and Turrigiano, G.G. (1998). BDNF has oppo- Mrsic-Flogel, T.D., Hofer, S.B., Ohki, K., Reid, R.C., Bonhoeffer, T., and site effects on the quantal amplitude of pyramidal neuron and interneuron Hu¨ bener, M. (2007). Homeostatic regulation of eye-specific responses in visual excitatory synapses. Neuron 21, 521–530. cortex during ocular dominance plasticity. Neuron 54, 961–972. Sarti, F., Schroeder, J., Aoto, J., and Chen, L. (2012). Conditional RARa Mu¨ ller, M., and Davis, G.W. (2012). Transsynaptic control of presynaptic Ca2+ knockout mice reveal acute requirement for retinoic acid and RARa in homeo- influx achieves homeostatic potentiation of neurotransmitter release. Curr. static plasticity. Front Mol Neurosci 5, 16. Biol. 22, 1102–1108. Schulz, D.J., Goaillard, J.-M., and Marder, E. (2006). Variable channel expres- Mu¨ ller, M., Pym, E.C.G., Tong, A., and Davis, G.W. (2011). Rab3-GAP controls sion in identified single and electrically coupled neurons in different animals. the progression of synaptic homeostasis at a late stage of vesicle release. Nat. Neurosci. 9, 356–362. Neuron 69, 749–762. Schulz, D.J., Goaillard, J.-M., and Marder, E.E. (2007). Quantitative expression Mu¨ ller, M., Liu, K.S.Y., Sigrist, S.J., and Davis, G.W. (2012). RIM controls profiling of identified neurons reveals cell-specific constraints on highly vari- homeostatic plasticity through modulation of the readily-releasable vesicle able levels of gene expression. Proc. Natl. Acad. Sci. USA 104, 13187–13191. pool. J. Neurosci. 32, 16574–16585. Seeburg, D.P., and Sheng, M. (2008). Activity-induced Polo-like kinase 2 is Muraro, N.I., Weston, A.J., Gerber, A.P., Luschnig, S., Moffat, K.G., and required for homeostatic plasticity of hippocampal neurons during epileptiform Baines, R.A. (2008). Pumilio binds para mRNA and requires Nanos and Brat activity. J. Neurosci. 28, 6583–6591. to regulate sodium current in Drosophila motoneurons. J. Neurosci. 28, 2099–2109. Seeburg, D.P., Feliu-Mojer, M., Gaiottino, J., Pak, D.T.S., and Sheng, M. (2008). Critical role of CDK5 and Polo-like kinase 2 in homeostatic synaptic Murdoch, J.D., and State, M.W. (2013). Recent developments in the genetics plasticity during elevated activity. Neuron 58, 571–583. of autism spectrum disorders. Curr. Opin. Genet. Dev. 23, 310–315. Shepherd, J.D., Rumbaugh, G., Wu, J., Chowdhury, S., Plath, N., Kuhl, D., Huganir, R.L., and Worley, P.F. (2006). Arc/Arg3.1 mediates homeostatic syn- Nerbonne, J.M., Gerber, B.R., Norris, A., and Burkhalter, A. (2008). Electrical aptic scaling of AMPA receptors. Neuron 52, 475–484. remodelling maintains firing properties in cortical pyramidal neurons lacking KCND2-encoded A-type K+ currents. J. Physiol. 586, 1565–1579. Soden, M.E., and Chen, L. (2010). Fragile X protein FMRP is required for homeostatic plasticity and regulation of synaptic strength by retinoic acid. O’Brien, R.J., Kamboj, S., Ehlers, M.D., Rosen, K.R., Fischbach, G.D., and J. Neurosci. 30, 16910–16921. Huganir, R.L. (1998). Activity-dependent modulation of synaptic AMPA recep- tor accumulation. Neuron 21, 1067–1078. Steinmetz, C.C., and Turrigiano, G.G. (2010). Tumor necrosis factor-a signaling maintains the ability of cortical synapses to express synaptic scaling. Paradis, S., Sweeney, S.T., and Davis, G.W. (2001). Homeostatic control of J. Neurosci. 30, 14685–14690. presynaptic release is triggered by postsynaptic membrane depolarization. Neuron 30, 737–749. Stelling, J., Sauer, U., Szallasi, Z., Doyle, F.J., 3rd, and Doyle, J. (2004). Robustness of cellular functions. Cell 118, 675–685. Penney, J., Tsurudome, K., Liao, E.H., Elazzouzi, F., Livingstone, M., Gonza- lez, M., Sonenberg, N., and Haghighi, A.P. (2012). TOR is required for the retro- Stellwagen, D., and Malenka, R.C. (2006). Synaptic scaling mediated by glial grade regulation of synaptic homeostasis at the Drosophila neuromuscular TNF-alpha. Nature 440, 1054–1059. junction. Neuron 74, 166–178. Stellwagen, D., Beattie, E.C., Seo, J.Y., and Malenka, R.C. (2005). Differential Petersen, S.A., Fetter, R.D., Noordermeer, J.N., Goodman, C.S., and regulation of AMPA receptor and GABA receptor trafficking by tumor necrosis DiAntonio, A. (1997). Genetic analysis of glutamate receptors in Drosophila factor-alpha. J. Neurosci. 25, 3219–3228. reveals a retrograde signal regulating presynaptic transmitter release. Neuron 19, 1237–1248. Sun, Q., and Turrigiano, G.G. (2011). PSD-95 and PSD-93 play critical but distinct roles in synaptic scaling up and down. J. Neurosci. 31, 6800–6808. Pilgram, G.S.K., Potikanond, S., van der Plas, M.C., Fradkin, L.G., and Noor- dermeer, J.N. (2011). The RhoGAP crossveinless-c interacts with Dystrophin Sutton, M.A., Wall, N.R., Aakalu, G.N., and Schuman, E.M. (2004). Regulation and is required for synaptic homeostasis at the Drosophila neuromuscular of dendritic protein synthesis by miniature synaptic events. Science 304, junction. J. Neurosci. 31, 492–500. 1979–1983.

Plomp, J.J., van Kempen, G.T., and Molenaar, P.C. (1992). Adaptation of Sutton, M.A., Ito, H.T., Cressy, P., Kempf, C., Woo, J.C., and Schuman, E.M. quantal content to decreased postsynaptic sensitivity at single endplates in (2006). Miniature neurotransmission stabilizes synaptic function via tonic alpha-bungarotoxin-treated rats. J. Physiol. 458, 487–499. suppression of local dendritic protein synthesis. Cell 125, 785–799.

Pratt, K.G., and Aizenman, C.D. (2007). Homeostatic regulation of intrinsic Sutton, M.A., Taylor, A.M., Ito, H.T., Pham, A., and Schuman, E.M. (2007). excitability and synaptic transmission in a developing visual circuit. Postsynaptic decoding of neural activity: eEF2 as a biochemical sensor J. Neurosci. 27, 8268–8277. coupling miniature synaptic transmission to local protein synthesis. Neuron 55, 648–661. Prinz, A.A., Bucher, D., and Marder, E. (2004). Similar network activity from disparate circuit parameters. Nat. Neurosci. 7, 1345–1352. Swensen, A.M., and Bean, B.P. (2005). Robustness of burst firing in dissoci- ated purkinje neurons with acute or long-term reductions in sodium conduc- Pulver, S.R., Bucher, D., Simon, D.J., and Marder, E. (2005). Constant ampli- tance. J. Neurosci. 25, 3509–3520. tude of postsynaptic responses for single presynaptic action potentials but not bursting input during growth of an identified neuromuscular junction in the lob- Tatavarty, V., Sun, Q., and Turrigiano, G.G. (2013). How to scale down post- ster, Homarus americanus. J. Neurobiol. 62, 47–61. synaptic strength. J. Neurosci. 33, 13179–13189.

Neuron 80, October 30, 2013 ª2013 Elsevier Inc. 727 Neuron Perspective

Temporal, S., Desai, M., Khorkova, O., Varghese, G., Dai, A., Schulz, D.J., and Van Wart, A., and Matthews, G. (2006). Impaired firing and cell-specific Golowasch, J. (2012). Neuromodulation independently determines correlated compensation in neurons lacking nav1.6 sodium channels. J. Neurosci. 26, channel expression and conductance levels in motor neurons of the stomato- 7172–7180. gastric ganglion. J. Neurophysiol. 107, 718–727. Wang, H.-L., Zhang, Z., Hintze, M., and Chen, L. (2011). Decrease in calcium Thiagarajan, T.C., Lindskog, M., and Tsien, R.W. (2005). Adaptation to synap- concentration triggers neuronal retinoic acid synthesis during homeostatic tic inactivity in hippocampal neurons. Neuron 47, 725–737. synaptic plasticity. J. Neurosci. 31, 17764–17771.

Tsurudome, K., Tsang, K., Liao, E.H., Ball, R., Penney, J., Yang, J.-S., Watt, A.J., and Desai, N.S. (2010). Homeostatic plasticity and STDP: keeping a Elazzouzi, F., He, T., Chishti, A., Lnenicka, G., et al. (2010). The Drosophila neuron’s cool in a fluctuating world. Front Synaptic Neurosci 2,5. miR-310 cluster negatively regulates synaptic strength at the neuromuscular junction. Neuron 68, 879–893. Weyhersmu¨ ller, A., Hallermann, S., Wagner, N., and Eilers, J. (2011). Rapid active zone remodeling during synaptic plasticity. J. Neurosci. 31, 6041–6052. Turrigiano, G. (2011). Too many cooks? Intrinsic and synaptic homeostatic mechanisms in cortical circuit refinement. Annu. Rev. Neurosci. 34, 89–103. Wojtowicz, J.M., and Atwood, H.L. (1983). Maintained depolarization of Turrigiano, G., Abbott, L.F., and Marder, E. (1994). Activity-dependent synaptic terminals facilitates nerve-evoked transmitter release at a crayfish changes in the intrinsic properties of cultured neurons. Science 264, 974–977. neuromuscular junction. J. Neurobiol. 14, 385–390.

Turrigiano, G.G., Leslie, K.R., Desai, N.S., Rutherford, L.C., and Nelson, S.B. Younger, M.A., Mu¨ ller, M., Tong, A., Pym, E.C., and Davis, G.W. (2013). A (1998). Activity-dependent scaling of quantal amplitude in neocortical neu- Presynaptic ENaC Channel Drives Homeostatic Plasticity. Neuron 79, 1183– rons. Nature 391, 892–896. 1196.

Tyagarajan, S.K., and Fritschy, J.M. (2010). GABA(A) receptors, gephyrin and Zhao, C., Dreosti, E., and Lagnado, L. (2011). Homeostatic synaptic plasticity homeostatic synaptic plasticity. J. Physiol. 588, 101–106. through changes in presynaptic calcium influx. J. Neurosci. 31, 7492–7496.

728 Neuron 80, October 30, 2013 ª2013 Elsevier Inc.