Mol Neurobiol DOI 10.1007/s12035-014-8768-8

The Role of Phosphoinositides in Synapse Function Yoshibumi Ueda

Received: 14 December 2013 / Accepted: 1 June 2014 # Springer Science+Business Media New York 2014

Abstract Since the discovery of phosphatidylinositol-3-kinase, scientific interest in the biological functions of phosphoinositides has greatly increased. Currently, seven phosphoinositides have been identified. These phosphoinositides are specifically localized to organelle membranes, their site of action. Phosphoinositides can regulate neuronal function by specifically recruiting downstream proteins that have phosphoinositidebinding domains. To date, it is well accepted that phosphoinositides play important roles in a broad spectrum of neuronal functions from regulating neural development to modulating synapse function. This review will provide an overview of the function and distribution of phosphoinositides at synapses. Keywords Phosphoinositide . Phosphatidylinositol 3,4,5-trisphosphate . Spinule . Synaptic plasticity . Förster resonance energy transfer . Fluorescence lifetime imaging Abbreviations PIP3 Phosphatidylinositol 3,4,5-trisphosphate PI(4,5)P2 Phosphatidylinositol 4,5-bisphosphate PI(3,4)P2 Phosphatidylinositol 3,4-bisphosphate PI(3,5)P2 Phosphatidylinositol 3,5-bisphosphate PI(4)P Phosphatidylinositol 4-phosphate PI(3)P Phosphatidylinositol 3-phosphate PI(5)P Phosphatidylinositol 5-phosphate PI3K Phosphatidylinositol 3-kinase PI4K Phosphatidylinositol 4-kinase PI(4)P5K Phosphatidylinositol 4-phosphate 5-kinase Y. Ueda (*) Department of Hematology and Immunology, Kanazawa Medical University, 1-1 Daigaku, Uchinada, Kahoku, Ishikawa 920-0293, Japan e-mail: [email protected] Y. Ueda e-mail: [email protected]

PI(5)P4K PI FRET AMPAR NMDAR CaMKII LTP LTD SHIP IP3 ER FAPP2 PTEN Fig4 MTMR OCRL INPP4A PLIP OSBP CERT

Phosphatidylinositol 5-phosphate 4-kinase Phosphatidylinositol Förster resonance energy transfer α-Amino-3-hydroxy-5-methyl-4isoxazolepropionic acid receptor N-methyl-D-aspartate receptor Ca2+/calmodulin-dependent protein kinase II Long-term potentiation Long-term depression Src homology 2-containing inositol 5′phosphatase Inositol 1,4,5-trisphosphate Endoplasmic reticulum PI(4)P adaptor protein 2 Phosphatase and tensin homolog Factor-induced gene Myotubularin-related protein 2 Oculocerebrorenal syndrome of Lowe Inositol polyphosphate 4-phosphatase 4a PTEN-like phosphatase Oxysterol-binding protein Ceremide transfer protein

Introduction The brain is comprised of billions of neurons that form intricately interconnected functional networks, which mediate higher order functions such as learning and emotional processing. Neurons are highly polarized cells with distinct compartments, including the cell body, axon, dendrites, and synapses, specialized junctions where inter-neuronal communication occurs. Neurons also have the unique ability to adapt their activity to different external inputs, a phenomenon known as plasticity. The cooperative interplay between

Mol Neurobiol

different signaling molecules and proteins is key for neuronal plasticity. How are complicated signals computed by neurons? One hypothesis is that compartmentalized signaling by cellular membranes is critically involved [1]. Neurons (as well as other cells) contain many membranous compartments such as the plasma membrane, mitochondria, endoplasmic reticulum (ER), Golgi apparatus and endosomes, which are all important sources of lipid membrane. Membranes serve as a place where signaling proteins can interact and initiate subsequent downstream signaling. Signaling proteins are recruited with high temporal precision to membranes by lipid second messengers. Phosphoinositides are one important group of lipid second messengers, characterized by an inositol ring and a glycerol backbone with two fatty acids (Fig. 1a). The diverse phosphorylation of D-3, D-4, and D-5 positions on the inositol ring gives rise to seven phosphoinositides species. Each phosphoinositide has a distinct localization within neurons and interacts with different types of proteins to enable compartmentalized signal transduction.

Emergence of Phosphoinositides Phospholipid signaling was revealed by the discovery of phospholipid turnover following acetylcholine stimulation in pigeon pancreas and guinea pig brain by Hokin and Hokin [23]. Subsequent studies revealed that phospholipid signaling is a universal signaling mechanism that can be evoked by a diverse range of growth factors and neurotransmitters. Furthermore, intense research lead to the identification of phosphatidylinositol 4,5bisphosphate (PI(4,5)P2) as the source of phospholipid turnover [24]. PI(4,5)P2 can be cleaved by phospholipase C (PLC) into two further products: inositol 1,4,5-trisphosphate (IP3) and diacylglycerol (DAG) [24]. IP3 increases Ca2+ through IP3 receptors on the ER. The increase in cytosolic Ca2+ leads to further Ca2+ release through ryanodine receptors on the ER. The Ca2+ increase induces Ca2+ entry through Orai channels at the plasma membrane [25]. IP3 signaling is involved in the regulation of a wide range of cell functions such as neural development, cell proliferation, fertilization, and sensory perception [26]. Studies from the Nishizuka group revealed that DAG works as an activator of protein kinase C (PKC) and is implicated in cellular functions such as cell differentiation and proliferation [27–29]. Thus, phosphoinositide turnover is important for a broad spectrum of signaling pathways and essential for the maintenance of cell function. A series of studies by the Martin and Holz groups in the early 1990s suggested that PI(4,5)P2 itself is required for secretion in chromaffin cells [30–32], which provided great contribution in the area of membrane trafficking. Additionally, in 1994, it was revealed that PI(4,5)P2 binds directly to the pleckstrin homology (PH) domain that a broad spectrum of signaling proteins occupy [33]. This evidence leads to the idea

that PI(4,5)P2 is not only the precursor of DAG and IP3, but can also function as a lipid second messenger. In 1988, the Cantley group reported that type I phosphatidylinositol kinase (PIK) specifically phosphorylates the D-3 ring position of the inositol moiety to generate a novel phospholipid, known as phosphatidylinositol 3-phosphate (PI(3)P) [6]. This report spearheaded the discovery of a new family of phosphoinositides, all of which are phosphorylated at the D-3 position on the inositol ring [6]. This included the isolation of phosphatidylinositol 3,4,5-trisphosphate (PIP3) in formyl–methionyl–leucyl–phenylalanine (fMLP)-activated neutrophils by the Sklar group (also in 1988) [34] and phosphatidylinositol 3,4-bisphosphate (PI(3,4)P2), again by the Cantley group in 1989 [35]. Additionally, in 1997, the Cantley group identified phosphatidylinositol 5-phosphate (PI(5)P), a phosphoinositide with phosphate on the D-5 position [36]. In the same year, the Ulug group discovered phosphatidylinositol 3,5-bisphosphate (PI(3,5)P2), a phosphoinositide with phosphate residues on both D-3 and D-5 positions [37]. Hence, the decade from 1988 to 1997 saw the discovery of many phosphoinositides and is viewed as a golden era for the field of lipid research.

Characteristic Properties and Species of Phosphoinositides Currently, seven phosphoinositides have been identified. The discovery of these different phosphoinositides variants highlights the diverse characteristics of phosphoinositides. Investigations (in mostly non-neuronal cells) have revealed insights into the relative abundance of different phosphoinositides (Fig. 1b; for an extensive review, see ref. [2]). The most abundant phosphoinositides are phosphatidylinositol 4phosphate (PI(4)P) and PI(4,5)P2 [2]. While the amount of PI(3,4)P2, PI(3,5)P2, PI(5)P and PIP3 are maintained at relatively low levels under basal conditions, stimulation can cause a transient increase in the amount, indicating that phosphoinositide levels are tightly regulated by neuronal activity [5, 38–40]. However, recent studies have revealed that an aberrant constitutive increase of phosphoinositides can induce neuronal dysfunction. For example, an abnormal elevation in PI(4,5)P2 levels at nerve terminals can increase the ratio of clathrin-coated vesicles in all synapses and reduce synaptic transduction [41]. The accumulation of PI(3,4)P2 induces excitotoxic neuronal death [17]. Additionally, phosphatase and tensin homolog (PTEN)-deficient mice, where PIP3/Akt signaling is constantly active can lead to Cowden syndrome, a condition characterized by hypertrophic and ectopic dendrites and axonal tracts, enhanced synaptogenesis, and behavioral changes including abnormal social interaction [42]. PI(3,5)P2 accumulation, on the other hand, causes Charcot–Marie–Tooth disease, a

Mol Neurobiol Fig. 1 The characteristics of phosphoinositides. a The general structure of phosphoinositides. b The relative amount of phosphoinositides. PI(4)P, PI(4,5)P2, PI(3,5)P2, and PI(3,4)P2 values are referred from ref. [2]. PI(3)P and PI(5)P values are calculated from refs. [3, 4]. The value for PIP3 is taken from ref. [5]. c PI is converted to PI(3)P by PI3K classes II and III (also known as Vps34) [6, 7], to PI(5)P by PIKfyve [8, 9], and to PI(4)P by PI4K [10]. PI is generated from PI(3)P by MTMR2 [11]. PI(3)P is converted to PI(3,5)P2 by PIKfyve [9]. PI(4)P is converted to PI(4,5)P2 by PI(4)P5K [10]. PI(5)P is converted to PI(4,5)P2 by PI(5)P4K [8, 13]. A PTENrelated 5-phosphatidylinositol phosphatase (PLIP) is a potential producer of PI from PI(5)P [14, 15]. PI(4,5)P2 is converted to PIP3 by PI3K [16] and PI(4)P by synaptojanin and OCRL [12]. PI(3,4)P2 is degraded to PI(3)P by INPP4a [17]. PIP3 is degraded to PI(4,5)P2 by PTEN [18] and to PI(3,4)P2 by SHIP [19], OCRL and group IV 5-ptase [20]. PI(3,5)P2 is converted to PI(5)P by MTMR2 [11, 21]. PI(3,5)P2 is degraded to PI(3)P by Fig4 [22]. PI is predominantly localized at endoplasmic reticulum, colored in yellow. PI(4)P is localized at Golgi apparatus in non-neuronal cells and synaptic vesicle in neurons colored in orange. PI(5)P, PI(3,5)P2, and PI(3)P function at the endosome highlighted in pink, while PI(3,4)P2, PI(4,5)P2, and PIP3 mainly work at plasma membrane colored in green

a

b

O

O O

O

HO P HO

6

1

2

3

OH OH 5

4

OH

PI PI(4)P PI(4,5)P2 PI(3)P PI(5)P PI(3,4)P2 PI(3,5)P3 PIP3

1000 50 50 2.5 0.5-2.5 0.4-0.8 0.4-0.8 0.05

c PI(5)P

MTMR2

PI(3,5)P2

Endoplasmic reticulum Golgi apparatus

PIKfyve

PLIP

Fig4

PIKfyve

Plasma membrane

PI3K classII, III

PI

Endosomal compartments

PI(3)P MTMR2

PI(5)P4K

PI4K

INPP4a

PI(4)P Synaptojanin OCRL

PI(3,4)P2 SHIP OCRL GroupIV 5-ptase

PI(4)P5K PI3K classI

PI(4,5)P2

PIP3 PTEN

DAG

severe hereditary motor and sensory neuropathy characterized by focally folded myelin sheaths and demyelination [21, 22]. These studies indicate that the cellular levels of phosphoinositides must be tightly regulated, as excessive amounts are strongly associated with disease. Therefore, understanding the enzymes involved in the synthesis and degradation of phosphoinositides is the key to homeostatically controlling the levels. The synthesis pathway and cellular localization of phosphoinositides were also investigated mainly using non-

IP3

neuronal cultured cells. The main precursor of these phosphoinositides is phosphatidylinositol (PI). PI is primarily synthesized from CDP-DAG and inositol by phosphatidylinositol synthase in the endoplasmic reticulum (ER) [43], and then transported to other organelles via lipid transfer proteins or vesicular trafficking (as shown in glucosylceramide recently [44]). PI is subjected to phosphorylation at each organelle where the specific enzymes responsible for phosphorylation at 3, 4, and/or 5 positions on the inositol ring are located. Several enzymes have been identified for each phosphoinositide

Mol Neurobiol

(Fig. 1c). The localization and function of different phosphoinositides is outlined in the following sections. PI(4)P PI(4)P is generated from PI by phosphatidylinositol 4-kinase (PI4K) [45–51] and is predominantly localized to the Golgi apparatus in non-neuronal cells [52, 53]. PI(4)P is important for a wide range of cell functions including trafficking from the Golgi apparatus to the endosome and non-vesicular lipid trafficking. Golgi apparatus to endosome trafficking is regulated by the binding of PI(4)P to clathrin adaptors such as epsinR [54] and AP-1 [55]. Additionally, PI(4)P works as an anchor for lipid-transfer proteins such as ceramide transfer protein (CERT) [56], oxysterol-binding protein (OSBP) [57], and PI(4)P adaptor protein 2 (FAPP2) [44]. For instance, CERT binds to the PH domain of PI(4)P, which promotes localization to the Golgi apparatus. The START domain in CERT can pull ceramide out of the ER and transport it to the Golgi apparatus presumably at ER-Golgi contact site [56]. Likewise, FAPP2 and OSBP regulates the transport of glucosylceramide from Golgi apparatus to the plasma membrane and cholesterol from the ER to Golgi apparatus, respectively. In neuronal cells, the function of PI(4)P on Golgi apparatus in the postsynaptic compartment remains unclear. However, at the presynaptic side, a report demonstrates that PI4K activity is concentrated at synaptic vesicles [45]. Additionally, synaptojanin, a protein involved in the dephosphorylation of PI(4,5)P2 to PI(4)P, is also present at the presynaptic side and plays an important role in the endocytosis of synaptic vesicles [53, 58]. These studies suggest that PI(4)P is localized at synaptic vesicles. PI(4)P is also produced by dephosphorylation from PI(4,5)P2 by oculocerebrorenal syndrome of Lowe (OCRL), which is responsible for the oculocerebrorenal syndrome. The dysregulation of PI(4,5)P2 metabolism to PI(4)P leads to several neurological diseases. OCRL mutation causes severe cognitive defects [59]. As for synaptojanin, PI(4,5)P2 metabolism is altered in Down syndrome model mice that have trisomic synj 1. Disomic treatment of synj 1 mice alleviated the aberrant PI(4,5)P2 metabolism and improved cognitive disabilities, suggesting that synaptojanin 1 may contribute to brain dysfunction and cognitive disabilities in Down syndrome [60]. PI(4,5)P2 PI(4)P is converted to PI(4,5)P2 by phosphatidylinositol 4phosphate 5-kinase (PI(4)P5K) [10]. PI(4)P5K isozymes consist of PI(4)P5Kα, β, and γ [10]. PI(4)P5Kγ is enriched in neurons [61]. PI(4,5)P2 production is also performed by PTEN from PIP3 [18] and by phosphatidylinositol 5-phosphate 4kinase (PI(5)P4K) from PI(5)P [8, 13]. Considering that PIP3 and PI(5)P are less abundant than PI(4,5)P2, PI(4)P could be

the main source of PI(4,5)P2. However a new function for PI(4)P has recently been identified, which differs from its role as a precursor for PI(4,5)P2 [62]. Using a specific probe, a study showed that PI(4,5)P2 was enriched at the plasma membrane of the soma, axons, dendrites, and spines of cultured hippocampal neurons [63]. However, it is considered that PI(4,5)P2 levels locally increase more [64]. The PIP2 can regulate endocytosis, exocytosis, and actin polymerization (described in later parts) by recruiting appropriate proteins through PH, ENTH, C2 domain, and so on as shown in Table 1. PIP3 PIP3 regulates a broad spectrum of neuronal functions, such as dendritic arborization [115], cell size [119], and axonal filopodia formation [120], as well as the function of synapses [121]. PIP3 is rarely present under basal conditions. However, a small amount of PIP3 is required for α-amino-3-hydroxy-5methyl-4-isoxazolepropionic acid receptor (AMPAR) clustering in spines [122]. PIP3 is generated from PI(4,5)P2 by class I phosphatidylinositol 3-kinase (PI3K) [16]. While PI3K is expressed throughout the neuron [123], the active form of PI3K (which is formed following the association with AMPAR is mainly localized to spines [124]. PIP3 is dephosphorylated by Src homology 2-containing inositol 5′-phosphatase (SHIP), OCRL, and group IV 5-ptase to PI(3,4)P2, and by PTEN to PI(4,5)P2. One report indicates that PTEN is mainly localized to the dendritic shaft [125]. The distinct localization of PTEN and active PI3K could determine where PIP3 accumulates in spines. Consistent with this idea, using a fluorescence lifetime-based PIP3 probe, our lab has shown that PIP3 is highly enriched in spines (compared to the dendritic shaft) of hippocampal CA1 pyramidal neurons [126]. Recently, PIP3 is also found at the presynaptic side and regulates syntaxin1A clustering and neurotransmitter release [123]. PI(5)P PI(5)P was the last member of the phosphoinositide family to be discovered and very little is known about its regulation and function. In non-neuronal cells, subcellular fractionation experiments revealed that PI(5)P is predominantly present in the plasma membrane [4]. PI(5)P is also observed in the smooth endoplasmic reticulum and Golgi apparatus to some extent [4]. PI(5)P can be generated by several different pathways. PIKfyve, the main enzyme involved in catalyzing PI(3)P to PI(3,5)P2, can also produce PI(5)P from PI [3, 8]. Another report shows that myotubularin-related protein 2 (MTMR2) can generate PI(5)P from PI(3,5)P2 [11, 21, 127, 128]. PI(5)P degradation is conducted by PI(5)P4K, an enzyme that converts PI(5)P to PI(4,5)P2. There are three isoforms of PI(5)P4K: PI(5)P4Kα, β, and γ [129]. PI(5)P4Kα is localized at the cytosol, while PI(5)P4Kβ is enriched in nucleus. The

Mol Neurobiol Table 1 A comprehensive list of neuronal phosphoinositide-binding proteins Name

Target

Binding domain

Function

Reference

EEA1 PIKfyve SNX13 Profilin

PI(3)P PI(3)P PI(3)P PI(4,5)P2

Excitatory synaptic transmission Degradation of Ca(V) 1.2 channels, endocytic cycling of AMPAR Neural development Spine morphology

[65] [66–68] [69] [70]

Cofilin N-WASP Gelsolin Talin Ezrin Radixin

PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2

FYVE FYVE PX Positive electrostatic potential 40 % surface of cofilin PH N and C terminus FERM FERM FERM

AMPAR trafficking Development of dendritic spines and synapses Ca2+ channel and NMDAR activities Molecular interactions underlying synaptic junctions Filopodial protrusion formation, growth cone guidance Filopodial protrusion formation, growth cone guidance

[71] [72] [73] [74] [75–77] [76, 77]

Moesin Filamin α-Actinin MARCKS Cortactin Spectrin PLCγ Synaptotagmin Doc2 Munc13 Rabphilin 3A CAPS Mint Piccolo/ aczonin RIM Syntaxin1

PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2

FERM Amino acid basic region Amino acid basic region Amino acid basic region Amino acid basic region PH PH C2 C2 C1 C2 C2 PTB C2

Filopodial protrusion formation, growth cone guidance Migration of neurons Spine morphology, NMDAR activity Maintenance of dendritic spine, LTP Dendritic spine formation, growth cone Spine morphogenesis, axonal transport of mitochondria Neuronal migration Ca2+ sensor for the exocytosis Ca2+ sensor for asynchronous exocytosis Priming factor for synaptic vesicle Priming factor for synaptic vesicle Priming factor for synaptic vesicle Priming factor for synaptic vesicle Priming factor for synaptic vesicle

[76, 77] [78] [79, 80] [81, 82] [83, 84] [85, 86] [87] [88] [89] [90] [91] [92] [93] [94]

PI(4,5)P2 PI(4,5)P2

Priming factor for synaptic vesicle SNARE protein for synapse fusion

[95, 96] [97]

Epsin AP180 CALM AP-2 (α,β2,μ2) Arf 6

PI(4,5)P2 PI(4,5)P2 PI(4,5)P2 PI(4,5)P2

Creating membrane curvature Clathrin-mediated Endocytosis Clathrin-mediated Endocytosis Clathrin-mediated Endocytosis

[98] [99] [100] [101]

PI(4,5)P2

Clathrin-mediated Endocytosis

[99]

Dynamin

PI(4,5)P2

C2 Electrostatics and hydrophobic A/ENTH A/ENTH A/ENTH Electrostatics and hydrophobic Electrostatics and hydrophobic PH

[102–105]

PLD Preso KCNQ

PI(4,5)P2 PI(4,5)P2 PI(4,5)P2

Scission of vesicles at presynaptic side, and AMPAR cycling at endocytic zone in spines Neurite outgrowth Spine morphology and dendritic outgrowth Regulate cell membrane potential and excitability in neurons

Formin ARNO SHC Tiam1

PI(3,5)P2 PIP3 PIP3 PIP3

Polarization of neurons Regulation of dendritic development Neuronal migration, Regulation of hippocampal synaptic plasticity Dendritic spine development

[110] [111] [87, 112] [113]

Sos Akt/PKB Protrudin

PIP3 PIP3, PI(3,4)P2 PIP3, PI(3,4)P2, PI(4,5)P2

LTP induction Dendritic arborization Neurite outgrowth

[114] [115] [116, 117]

PH FERM Electrostatics and hydrophobic PTEN domain PH PTB PH PH PH FYVE

[106] [107, 108] [109]

For a comprehensive review on the proteins that function in other cell types (as well as additional information about neuronal cells) please refer to refs. [53, 118]

Mol Neurobiol

localization of PI(5)P4Kγ is unclear. The enzymatic activity of PI(5)P4Kα is much higher than PI(5)P4Kβ [129]. Additionally, PTEN-related 5-phosphatidylinositol phosphatase (PLIP) is a potential negative regulator of PI(5)P, causing degradation to PI [15]. Currently, several PI(5)P-specific binding domains have been identified. Dok-1, Dok-2, Dok-4, and Dok-5 selectively bind to PI(5)P through tandem domains of phosphotyrosine-binding (PTB) and pleckstrin homology (PH) domains [130, 131]. ING1, ING2, and ACF are nucleus-localized proteins, which are also associated with PI(5)P via the plant homeodomain (PHD) [8]. A report also indicates that ING2 regulates p53 acetylation via PI(5)P association [132]. In neurons, MTMR2 is enriched in spines. Knockdown of MTMR2 leads to a decrease in EEA1 clusters in dendrites. Hence, PI(5)P is presumably present at early endosomes [133]. In another report, PI(5)P4K was mostly present in vesicular compartments of the somatic cytoplasm of hippocampal neurons [13]. PI(3,4)P2 PI(3,4)P2 is generated from PIP3 upon activation of SHIP, OCRL, and group IV 5-ptase [19, 20] and dephosphorylated by inositol polyphosphate 4-phosphatases 4a (INPP4a) to PI(3)P [17]. INPP4a is deficient in Weeble mutant mice that exhibit significant neuronal loss in the cerebellum and hippocampal CA1 field, indicating that abnormal PI(3,4)P2 accumulation could cause neuronal dysfunctions [134]. Additionally, another study shows that PI(3,4)P2 regulates the number of Nmethyl-D-aspartate receptors (NMDARs) at spines, indicating that PI(3,4)P2 is enriched around NMDAR in spines [17]. There are two isoforms of SHIP: SHIP1 and SHIP2. The importance in SHIP is extensively studied in immunology. For example, SHIP1 KO mice leads to a severe myeloproliferative disorder and impaired NK cell function [135]. A study using migrating MDCK cells demonstrated that PI(3,4)P2 levels is higher in the front than in the rear of the cell, presumably mediating SHIP activation [136]. PI(3,4)P2 binds to downstream-signaling proteins including Akt and PDK1 via the PH domain, and TAPP1 and TAPP2 via the tandem PH domain containing protein (TAPP) domain. Currently, there are few reports about SHIP in neurons. However, SHIP2 downregulates PIP3 signaling by degrading PIP3 to PI(3,4)P2 upon nerve growth factor-induced neuritogenesis in PC12 cells [19]. PI(3)P PI(3)P regulates a broad range of cellular functions such as endomembrane membrane fusion [137], vesicular trafficking [138], and autophagy [139]. PI(3)P is produced from several pathways. PI3K class II produces PI(3)P at the plasma membrane and clathrin-coated vesicles in response to hormones [140]. PI3K class III, also known as Vps34, produces PI(3)P

at endomembranes [141]. Additionally, INPP4a dephosphorylates PI(3,4)P2, generating PI(3)P [17]. Furthermore, PI(3)P is produced by FIG4 from PI(3,5)P2 [22]. PI(3)P accomplishes its functions by recruiting downstream signaling proteins through different binding modules including the FYVE and PX domains. For instance, early endosome antigen-1 (EEA1) is localized at early endosomes through a FYVE domain and also recruits Rab5, leading to the fusion of early and late endosomes [137]. In hippocampal neurons, PI(3)P is primarily localized to dendrites and axons and partially to spines [142]. PI(3,5)P2 PI(3,5)P2, in combination with PI(3)P, regulates multiple aspects of endo-membrane trafficking in neuronal cells [142]. PI(3,5)P2 is generated from PI(3)P by PIKfyve and degraded to PI(3)P by FIG4 [9]. Interestingly, it is known that FIG4 and PIKfyve interact indirectly via an associated regulator of PIKfyve (ArPIKfyve) [143]. The complex is attached on the endosome through a EYVE domain and functions to regulate PI(3,5)P2 levels. In non-neuronal cells, PI(3,5)P2 is widely known for its role in regulating membrane trafficking, vacuole formation, and the function of endosomes and lysosomes [143, 144]. In neurons, PIKfyve regulates NMDA-induced voltagegated Ca2+ channel internalization [66] and postsynaptic trafficking of AMPAR to the plasma membrane [67].

Phosphoinositide-Binding Proteins in Neurons In the past, PI(4,5)P2 was considered just as a precursor for DAG and IP3. However, it was revealed that PI(4,5)P2 can bind to the pleckstrin homology (PH) domain of other proteins [33]. Since then, several new functions for PI(4,5)P2 have emerged. Studies have indicated that PI(4,5)P2 can directly interact with other proteins via the PH domain, to recruit specific signaling proteins to the membrane and enable control of downstream signaling pathways. Currently, it is well known that a variety of signaling proteins contain a phosphoinositidebinding domain. Table 1 shows a list of the signaling proteins reported to function in neurons. However, further intense examination is still required to determine how the interaction between phosphoinositides and the binding domains regulates neuronal function. For comprehensive reviews on phosphoinositide-binding proteins, please see refs. [53, 118].

Techniques for the Detection of Cellular Phosphoinositides Early studies utilised biochemical methods to measure cellular levels of phosphoinositides. Typical experiments involved the metabolic labeling of phosphoinositides with radioisotopes

Mol Neurobiol

(e.g., 14C, 3H, and 32P), followed by tissue homogenization and then extraction. In some cases, the lipids were subjected to the deacetylation of phosphatidylglycerol structures or the methylation of phosphates on the inositol ring after extraction for better separation of structurally similar lipids using highpressure lipid chromatography (HPLC; even without radioisotopes [145]). Recent advances in liquid chromatography and mass spectrometry analysis have greatly contributed to our ability to analyze lipids [5, 146–148]. These methods allow us to measure the absolute values of phospholipids and to obtain the lipid profile depending on the difference in fatty acid composition. When combined with fractionation techniques, this method enables the measurement of phosphoinositide levels at specific organelle membranes [148]. One report using synaptosome fractions showed that PI(4,5)P2 levels can be measured as detergent soluble and insoluble membranes [149]. However, the main drawback of these methods is that one cannot obtain the time course or single-cell profile of phosphoinositide dynamics. In addition, when subcellular distribution of phosphoinositides (Fig. 2) is

Fig. 2 Subspine distribution of phosphoinositides. PI(3)P is enriched in spines [142]. PI(3,5)P2 and PI(5)P is localized at endosomal compartments [133]. Additionally PI(3,5)P2 is involved in GluR1 synaptic vesicle trafficking [67]. PI(4)P is localized at synaptic vesicles because PI4K activity is detected [45]. PI(4,5)P2 is localized at the plasma membrane of the soma, axons, dendrites, and spines [63]. PI(3,4)P2 regulates the number of NMDARs at the synapse, indicating that PI(3,4)P2 is enriched around NMDAR in spines [17]. PIP3 regulates AMPAR clustering [122] and is found at presynaptic side, regulating SNARE [123]. Additionally, synaptotagmin is able to associate with PIP3 at basal condition [150]

examined by fractionation, the contamination from one membrane fraction to another should be taken into consideration. In order to observe phosphoinositides in single cells, fluorescent imaging tools were developed. Since the late 1990s, genetically encoded fluorescent protein-linked lipid-binding domains have been engineered to enable the specific study of phosphoinositide dynamics and functions in live cells. To examine PIP3 activity, PIP3-binding domains derived from Btk [151] and GRP1 [152] were used as indicators for PIP3. Fluorescent probes can be generated by creating fusion proteins, made up of PIP3-binding domains and fluorescent proteins [such as green fluorescent protein (GFP)]. These probes can then be expressed in living cells. The probes are evenly distributed in the cytosol. In response to different conditions or stimuli, the probes translocate to the cellular membrane where PIP3 is generated. Based on the extent of probe accumulation, we can identify the level of PIP3 production. The same strategy has been applied to PI(4,5)P2 using a PH domain from PLCγ [63, 153]. However, several considerations must be taken into account when interpreting imaging

Endosome Synaptic vesicle

Post synatptic density PI(4)P PI(3,5)P2 PIP3 PI(4,5)P2 PI(3,4)P2 PI(3)P PI(5)P AMPAR NMDAR Glutamate

Early endosome

Late endosome

Mol Neurobiol

experiments. As phosphoinositide-imaging experiments are typically performed in live cells, any alterations in cell shape or encounters with uneven membrane structures (e.g., membrane ruffles) can significantly affect the fluorescence intensity measurements, potentially leading to the generation of artifacts. Moreover, it can often be difficult to distinguish which membranous compartments the fluorescent fusion proteins translocated to. In order to overcome these limitations, we have developed Förster resonance energy transfer (FRET)based lipid probes for PIP3 [154] and DAG [155]. Our PIP3 FRET probe was constructed as follows (Fig. 3a): to enable specific PIP3 binding, a PH domain from GRP1 was introduced between cyan fluorescent protein (CFP) and a yellow fluorescent protein (YFP) variant, through rigid α-helical linkers consisting of repeated EAAAR sequences. We introduced a single di-glycine motif within one of the rigid linkers, which acts as a hinge and provided flexibility to the probe. A CAAX sequence from N-Ras was also attached to the probe to

Fig. 3 PIP3 probe and fluorescence lifetime imaging of PIP3 in spines. a PIP3 is produced PI3K, and then PIP3 on the plasma membrane can interact with the PH domain of the probe. This interaction induces a conformation change in the probe leading to an increase in FRET. b Images of fluorescence lifetimebased PIP3 probe (FLIMPA3) and FLIMPA mutant (mut), which is not able to bind to PIP3, were expressed in CA1 pyramidal neurons in hippocampal slices. The FLIMPA3 probe demonstrated that PIP3 is enriched at spines (compared to dendrites), but this enrichment was not observed in neurons expressing the FLIMPA mutants

assist plasma membrane targeting. The binding of PIP3 to the PH domain leads to a substantial change in the conformation of the probe. This flip-flop-type conformational change results in increased FRET from CFP to YFP and allows stable observation of PIP3 by dual-emission ratiometric imaging. Ratiometric imaging can cancel out the artifact derived from fluorescent intensity changes caused by membrane ruffling or changes in cell shape. There are two main strengths to this probe. Firstly, by using specific membrane target sequences, it enables the accurate examination of PIP3 dynamics at the membrane of interest. Secondly, the binding domain can be easily modified to enable the study of other phosphoinositides such as PI(4,5)P2, PI(4)P, and PI(3,4)P2 (as reported by Matsuda group) [19, 136, 156]. Additionally, for studies in thick tissue such as hippocampal slices or intact brains, we specifically engineered a fluorescence lifetime-based probe, which only considers the fluorescence lifetime of GFP and effectively eliminates any artifact caused by wavelength-

a PH YFP

CFP

Gly-Gly

αhelix (EAAAR)7 PIP3

PI3K activation Membrane Membrane target domain

b

High

Low PIP3

Mol Neurobiol

dependent absorption that is often observed in thick tissue preparations [126].

Neuronal Functions of Phosphoinositides Excitatory synapses consist of two main components, a presynaptic terminal and a postsynaptic apposition. The presynaptic terminal is characterized by the presence of vesicles containing neurotransmitters [157]. The postsynaptic side is defined by the presence of micron-sized membrane protrusions, commonly known as spines. Within spines, there are several organelles and structures including the protein-rich structure called the postsynaptic density (PSD), where multiple interactions between different proteins, the smooth endoplasmic reticulum (SER), spine apparatus derived from SER, early and recycling endosomes, and lysosomes [157–159]. In addition to being the junctions of neurotransmission, synapses are also the elementary unit that can undergo synaptic plasticity, a phenomenon that underlies learning and memory [160]. Phosphoinositides are critically involved in regulating a wide range of synaptic functions including synaptic vesicle trafficking at the presynaptic side, and modulation of spine morphology and receptor exocytosis/endocytosis on the postsynaptic side [17, 126, 161, 162]. In this section, I will review the roles of phosphoinositides on modulating presynaptic and postsynaptic functions and plasticity mechanisms (for review, see ref. [58] for a comprehensive review on this topic).

Presynaptic Side The process of neurotransmitter release at nerve terminals consists of a series of exquisitely orchestrated steps [123, 163]. Firstly, synaptic vesicles need to be “primed” for neurotransmission. For this to occur, the synaptic vesicles bud from plasma membrane and then filled with neurotransmitter. The vesicles then translocate and dock at the active zone of the terminal, a place where proteins required for priming (such as Munc13, CAPS, and RIM) are present [92, 164, 165]. These docked and primed vesicles are at the penultimate stage of neurotransmitter release. However, neurotransmission can only occur upon arrival of an action potential, the ultimate trigger for vesicle fusion. Arrival of an action potential at the nerve terminal leads to the activation of voltage-gated Ca2+ channels, which causes a local and transient increase in intraterminal Ca2+. This elevation in Ca2+, subsequently, triggers the fusion of vesicles to the plasma membrane via the SNARE complex, resulting in neurotransmitter release. After fusion, the plasma membrane undergoes clathrin-mediated endocytosis, enabling the synaptic vesicle to be retrieved. (Since this review mainly focuses on phosphoinositide function in

synapses, please see refs. [64, 88] for further information on synaptic vesicle recycling.) Recent findings have demonstrated that the phosphoinositides have important functions in regulating presynaptic neurotransmission. PI(4,5)P2 is implicated in multiple stages of synaptic vesicle cycling including priming, fusion, and endocytosis [64, 166]. Priming requires the participation of a variety of proteins including Munc13, RIM, and CAPS [167]. Other proteins, such as ELKS and Piccolo/Bassoon, RIM-binding proteins, and Liprins, form a complex, which functions as a scaffold for synaptic vesicles at the active zone [168]. Interestingly, several proteins related to priming have PI(4,5)P2-binding domains (such as C2 and PTB domains) as shown in Table 1, potentially indicating that these proteins are regulated by PI(4,5)P2. In particular, the Südhof group showed that PI(4,5)P2 binds to the C2B domain of Munc13 in a Ca2+-dependent manner [169]. Introducing mutations into PI(4,5)P2 affected its binding ability to the C2B domain of Munc13, which lead to changes in neurotransmitter release. This indicated that PI(4,5)P2 controls synaptic vesicle exocytosis through the C2B domain of Munc13. In sum, this evidence suggests that PI(4,5)P2 plays an important role in the priming step through the binding of proteins necessary for priming. PI(4,5)P2 is also important for vesicle fusion. During fusion, PI(4,5)P2 works cooperatively with the synaptic vesicle Ca2+ sensor, synaptotagmin to regulate synaptic vesicle fusion. While the C2B domain of synaptotagmin binds to PIP3 in the absence of Ca2+, it subsequently binds to PI(4,5)P2 in the presence of Ca2+, suggesting that a lipid interaction switch occurs during depolarization [150]. The increase in Ca2+ triggers the association between the C2 domain and PI(4,5)P2. This leads to the insertion of the C2 domain into the plasma membrane, which promotes plasma membrane curvature and, subsequently, enhances the efficiency of vesicle fusion [99, 162]. After neurotransmitter release, synaptojanin dephosphorylates PI(4,5)P2 to generate PI(4)P during vesicle endocytosis. This reaction leads to the detachment of clathrin from the vesicle [41]. This evidence indicates that synaptic vesicles are enriched in PI(4)P and contain less PI(4,5)P2. The idea is supported by a study examining the subcellular distribution of PI(4,5)P2 in cultured hippocampal neurons using a PH domain–GFP probe [63]. Interestingly, a study revealed that FM4-64, a synaptic vesicle marker, and a PI(4,5)P2 probe were mutually exclusive at presynaptic boutons, suggesting a lack of significant amounts of PI(4,5)P2 on synaptic vesicles at quiescent synapses. Furthermore, the Chapman group recently showed using in vitro fusion assays with reconstituted SNARE proteins that while PI(4,5)P2 in vesicle membranes is not needed for vesicle fusion, PI(4,5)P2 in the target vesicle membrane is required [170]. Taken together, PI(4,5)P2 plays many important roles in the multistep process of synaptic vesicle cycling.

Mol Neurobiol

Postsynaptic Side Spines are tiny membranous structures located on dendritic shafts. Spines can exhibit both functional and morphological plasticity in response to presynaptic stimulation and therefore serves as a flexible component within an otherwise fairly rigid neuronal network of the adult brain. Long-term potentiation (LTP) of synaptic transmission is thought to be an important mechanistic model that underlies memory formation [160]. The molecular mechanism for LTP at CA3–CA1 synapses can be explained as follows. The administration of high-frequency electrical stimulation to CA3 presynaptic neurons causes the release of neurotransmitters (including glutamate) from nerve terminals and, subsequently, leads to membrane depolarization of postsynaptic CA1 neurons. These events trigger the removal of Mg2+ from the channel pore of NMDAR, leading to NMDAR activation and further influx of Ca2+ through NMDAR. The elevation in Ca2+ can recruit AMPARs to the surface of spines from either extrasynaptic regions or intracellular compartments by the activation of signaling molecules, including Ca2+/calmodulin-dependent protein kinase II (CaMKII) [171]. The insertion of AMPARs at the postsynaptic membrane results in an increase in electrical transmission. Conversely, the application of low-frequency electrical stimulation decreases the electrical transmission, known as longterm depression (LTD). LTD causes the internalization of AMPARs from the membrane surface, a process mediated by calcineurin [172]. LTP and LTD are widely considered as possible models for the molecular events that underlie learning and memory [173]. Spines come in a variety of shapes and are typically classified into the following categories: mushroom, stubby, thin and filopodia [174]. Interestingly, the structure of spines is often deeply associated with synaptic plasticity [174] as well as neural diseases such as mental retardation [175]. Electron microscopy has revealed the presence of even finer structures on spines, including filopodia like protrusions, spinules, and spines perforations, all of which can dynamically change their structure in response to theta burst stimulation [176] and glutamate application [177]. Additionally, spine size and AMPAR current are positively correlated, indicating a strong relationship between structural and functional changes [178]. To date, phosphoinositides have been implicated in a broad spectrum of spine functions including current transmission, spine morphology and size, and trafficking of AMPAR and NMDAR containing vesicles.

Functional Plasticity The phosphoinositides have been implicated in regulating functional plasticity, with PIP3 being one of the most commonly studied phosphoinositides. The Francesconi group

showed that theta burst-induced LTP in hippocampal CA1 pyramidal neurons was abolished in the presence of a PI3K inhibitor [121]. In addition, the Wang group showed that PI3K is directly associated with AMPAR when glycine-induced specific activation of NMDAR occurs. This suggests that PI3K activates and produces PIP3 at close spatial proximity to AMPAR [124]. Furthermore, the Esteban group reported that PIP3 in spines is crucial for maintaining AMPAR clustering [122]. Their findings indicated that a decrease of PIP3 in spines causes the movement of AMPAR from the postsynaptic density towards the perisynaptic membrane within the spine, which leads to synaptic depression. In another paper, they investigated the effect of PTEN on LTP and LTD in CA1 pyramidal neuronal cells [179]. PTEN is a protein that can decrease PIP3 levels. Interestingly, overexpression of PTEN in neurons abolished basal AMPA current. Next, the effect of PTEN inhibitor on LTD was examined. Incubation of slices with a PTEN inhibitor abolished electrically induced LTD. In summary, the regulation of PIP3 levels by PI3K and PTEN signaling plays an important role in modulating LTP and LTD by influencing the movement of AMPAR in spines. PI(3,5)P2 is also involved in the trafficking of AMPAR to the plasma membrane on the postsynaptic side [67]. In hippocampal neurons, serum- and glucocorticoid-inducible kinase 3 (SGK3) mRNA is upregulated after NMDAR activation. Subsequently, SGK3 can phosphorylate PIKfyve at serine 318 to activate it. Activated PIKfyve then produces PI(3,5)P2, which activates Rab11 located on AMPARcontaining vesicles, leading to the trafficking of vesicles to the surface of postsynaptic spines. It was also shown in this report that injection of PI(3,5)P2 into Xenopus oocytes expressing GluA1 could amplify the GluA1 current. PI(3,4)P2 also plays an important role in functional plasticity. INPP4A is an enzyme responsible for degrading PI(3,4)P2 to PI(3)P. The accumulation of this lipid induces glutamate excitotoxicity in the central nervous system and exhibits involuntary movement and mortality in INPP4A−/ −mice [17]. This study also showed that PI(3,4)P2, but not PIP3 and PI(4,5)P2, can massively increase glutamateinduced excitotoxicity through the accumulation of NMDARs at synapses. There are several reports that PI(4,5)P2 is involved in LTD in hippocampal neurons. The Dell’Acqua group showed that protein kinase A kinase-anchoring protein (AKAP)79/150 plays a pivotal role in NMDAR-dependent LTD [153, 180]. AKAP79/150 is targeted to spines through an N-terminal basic region that binds PI(4,5)P2, F-actin, and cadherin and forms a complex with PKA and calcineurin (CaN)/protein phosphatase 2B (PP2B) at static state. NMDA stimulation activates PLC, which is followed by degradation of PI(4,5)P2. A decrease in PI(4,5)P2 removes AKAP79/150 and PKA from spines, leading to the inability of AMPARs to be phosphorylated by PKA, resulting in LTD. This data

Mol Neurobiol

suggests that a decrease in PI(4,5)P2 in spines induces LTD. In contrast, the Kanaho group showed that an NMDA-triggered increase in PI(4,5)P2 can induce LTD [181]. Ca2+ elevation by NMDAR activation activates protein phosphatase 1 (PP1) and calcineurin, causing dephosphorylation of PI(4)P5Kγ661, a splice variant of PI(4)P5Kγ at spines. The dephosphorylated PIP5Kγ661 increases the affinity to AP-2. The association with AP-2 activates PIP5Kγ661, which then leads to PI(4,5)P2 production. The newly synthesized PI(4,5)P2 then recruits endocytic components, which result in the endocytosis of AMPAR. The difference in PI(4,5)P2 dynamics during LTD between the two studies can be explained by a difference in the localization of PI(4,5)P2. At first, PI(4,5)P2 breakdown by PLC after NMDAR activation occurs at the postsynaptic density. This removes AKAP79/150 and PKA from spines, leading to AMPAR dephosphorylation. The dephosphorylated AMPAR then moves to the active endocytic zone, where PI(4)P5Kγ661 produces PI(4,5)P2.

Structural Plasticity/Spine Morphology Spine morphology is regulated by actin polymerization and depolymerization [174, 182, 183]. Considering that PI(4,5)P2 binds to almost all of the actin-binding proteins, it is reasonable to assume that PI(4,5)P2 may play an important role in regulating spine morphology. The Inoue group investigated the effect of PI(4,5)P2 levels on actin dynamics using a brandnew technique, rapamycin-induced protein dimerization in culture cells [184]. When PI(4,5)P2 was increased by rapamycin-induced recruitment of PI(4)P5K to the plasma membrane, COS-7, HeLa, and HEK293 cells formed bundles of motile actin filaments known as actin comets. This evidence directly suggests that PI(4,5)P2 regulates actin dynamics. Additionally, recent studies showed that Preso, which interacts with PI(4,5)P2 through the FERM domain can regulate spine morphology and dendritic outgrowth [107, 108]. Furthermore, several studies showed that PI(4,5)P2-binding proteins can regulate spine morphology and number [70, 79, 81, 83, 86]. However, direct evidence that PI(4,5)P2 exclusively regulates spine morphology through the PI(4,5)P2binding domains is currently limited, because the PI(4,5)P2binding regions of cofilin, profilin, gelsolin, etc., are often shared with the actin-binding region [185–187]. This characteristic makes it more challenging to studying the implications of PI(4,5)P2 on spine morphology. As for other phosphoinositides, a loss of PI(3)P in hippocampal neurons from conditional VPS34 deletion mice was associated with a loss of spines, and ultimately neurodegeneration [142]. The development of the glutamate uncaging method in 2004 has made a huge advancement to our understanding of the molecular mechanisms that underlie structural LTP (sLTP)

[178]. After glutamate is locally applied to a spine of interest by photolysis of caged glutamate, the spine becomes around three to four times bigger than the basal spine size, approximately 1–2 min after stimulation. The spine size then reaches a plateau, before shrinking slightly, but still maintains a larger size (compared to the size before stimulation) even after several hours [178]. The difference in spine size before and after glutamate stimulation is often interpreted as a potential mechanism that may underlie learning and memory. To date, several key players for the regulation of sLTP have been revealed as follows [188]. Uncaged glutamate induces an increase in Ca2+ concentration through NMDAR [178]. The Ca2+ increase activates calmodulin/CaMKII [189], leading to the activation of small GTPases such as Ras, cdc42, and RhoA [190]. These signals, subsequently, converged in the LIM kinase/cofilin pathway to modify spine size through actin polymerization/ depolymerization [182]. Recently, we investigated the function of PIP3 in regulating the spine morphology of CA1 pyramidal neurons in the hippocampus [126]. We discovered that glutamate uncaging induced not only an enlargement of spines but also the formation of small filopodia-like protrusions, termed spinules. Interestingly, spinules were observed during the first 2 min after stimulation on 50 % of the spines. The number of spinules increased in the presence of a PTEN inhibitor, which increased PIP3. Conversely, the number of spinules decreased following treatment with a PI3K inhibitor. Together, these results indicate that PIP3 regulates spinule formation. In order to visualize PIP3 dynamics in spines, we developed a fluorescence lifetime-based PIP3 probe, FLIMPA3. The probe was based on a ratiometric PIP3 FRET probe, but the donor CFP and acceptor YFP molecules were exchanged to mEGFP and sREACh, respectively, to obtain the fluorescence lifetime signal. When FLIMPA3 was expressed in CA1 pyramidal neurons in hippocampal slices and observed using two-photon microscopy, it was found that PIP3 was more highly concentrated in spines, compared to dendrites (Fig. 3b). There are reports in which active PI3K, which is formed just after the association with AMPAR, mainly resides in spines, while PTEN is localized to the dendritic shaft of CA1 neuronal pyramidal cells [124, 125]. Therefore, PIP3 accumulation is determined by the localization of PTEN and active PI3K. Next, we observed PIP3 dynamics during sLTP. After glutamate uncaging, basal PIP3 in the spine was reduced inversely to the enlargement of the spine. Interestingly, the reduction of PIP3 after stimulation was highly correlated with relative PIP3 enrichment in spines (compared to the dendritic shaft) before the stimulation. Therefore, the reduction in PIP3 is primarily due to the addition of membrane from the dendritic shaft. Furthermore, while PIP3 decreases globally in spines during stimulation, we observed a specific accumulation of PIP3 in spinules. Overall, this data suggest that PIP3 regulates spinule formation on spines subjected to sLTP induction. Additionally, we found that PIP3 accumulates in spinules during sLTP.

Mol Neurobiol

There are several studies investigating the properties of spinules using electron microscopy (EM). EM can highlight finer structures such as spinules and perforations on spines. Studies have shown that the number of spinules increases in response to theta burst stimulation [176], local glutamate stimulation [177], and high potassium application [191]. The length of spinules typically increase in response to exogenously applied glutamate [177]. EM studies have demonstrated that spinules could also be trans-synaptically endocytosed as separate vesicles by presynaptic terminals from the postsynaptic side [192]. Therefore, the spinules may serve as a mechanism for retrograde signaling or may aid postsynaptic membrane remodeling by removing excess membrane (Fig. 4) [192]. Alternatively, the McKinney group showed that the glutamate-induced extension of spines can lead to the formation of new synapses [177]. Compared to its postsynaptic role, the function of PIP3 at presynaptic terminals remains enigmatic. However, a recent Drosophila study showed that presynaptic PIP3 interacts with syntaxin 1A, a protein essential for vesicle fusion [123]. Modulation of PIP3 levels can change syntaxin 1A clustering and, subsequently, regulate neurotransmitter release. Furthermore, synaptojanin, another synaptic vesicle protein, was demonstrated to have higher affinity to PIP3 than PIP2 under basal conditions. Based on this evidence, the accumulation of PIP3 in spinules may influence both presynaptic and postsynaptic function. Fig. 4 Summary of the potential physiological significance of PIP3 accumulation in spinules. This figure outlines some of the possible biological functions of PIP3 in spinules. The stimulation of spines (a) can lead to the generation of spinules (b) on spines. Currently, there are three proposed roles for spinules. c-1 PIP3 signaling may occur at spinules to enable new synapses to form with functional presynaptic buttons [177]. c-2 PIP3 can be sent to the presynaptic site and may act as a messenger molecule for retrograde signaling [192]. c-3 The extended spinules may just be by-products of the process of spine reorganization [192]

Concluding Remarks Since the discovery of phosphoinositides, our knowledge of the function of different phosphoinositides species has greatly expanded. While the function of PI(4,5)P2 and PIP3 at synapses has been intensively studied, other synaptic phospholipids have been largely ignored. Interestingly, PI(4,5)P2, PIP2, PI(3,4)P2, and PI(3,5)P2 are involved in the regulation of proteins such as AMPAR and NMDAR, key components underlying synaptic plasticity. How do different phosphoinositide species and signaling pathways influence the dynamics and function of different proteins? One plausible explanation is the compartmentalized signaling of each phosphoinositide. In order to address the question further, information about subspine phosphoinositide dynamics is required. However, as the size of spines is very small, just ~1 μm, conventional light microscopy may not provide sufficient resolution to reliably study subspine activity as the resolution limit attained by light microscopy is close to the size of a spine. Super resolution fluorescent microscopy, such as STORM/PALM and STED, can provide a more detailed view of phosphoinositide function [193–196]. Another future direction is the relevance of phosphoinositides in disease states. Lithium has been an effective drug to improve bipolar disorder [197, 198]. The drug affects inositol monophosphatase (c)-1. New synapse formation

(c)-2. Retrograde signaling

PIP3 Low

(a)

High

(b)

Postsynaptic density Structural LTP stimulation

(c)-3. Spine reorganization

Mol Neurobiol

and GSK3β, downstream signaling proteins of PIP3 [198, 199]. PTEN mutations cause Cowden disease, which is often accompanied by mental retardation, and are associated with an autism spectrum disorder [200]. The proteins related to PI(4,5)P2 generation and degradation, such as PI4K, PI(4)P5K, and synaptojanin, are related to bipolar disorder [201]. Synaptojanin 1 is also related to Alzheimer’s disease. Synj1 haplo-insufficient mice is protective against amyloidβ [202, 203] and induces its clearance [204]. Furthermore, there are several reports in which synj1 overexpression could be deleterious in Down syndrome [60, 205–208]. Acknowledgements I am grateful to L. Yu and F. Hullin-Matsuda for comments on the manuscript. I also thank the members of Hayashi’s laboratory for their support, critical reading of the manuscript and discussion. This work was supported by a Grant-in-Aid for Young Scientists (B) and the RIKEN Special Postdoctoral Researcher Program.

References 1. Bivona TG, Perez De Castro I, Ahearn IM, Grana TM, Chiu VK, Lockyer PJ, Cullen PJ, Pellicer A, Cox AD, Philips MR (2003) Phospholipase Cgamma activates Ras on the Golgi apparatus by means of RasGRP1. Nature 424(6949):694–698 2. Lemmon MA (2008) Membrane recognition by phospholipidbinding domains. Nat Rev Mol Cell Biol 9(2):99–111 3. Shisheva A (2001) PIKfyve: the road to PtdIns 5-P and PtdIns 3,5P(2). Cell Biol Int 25(12):1201–1206 4. Sarkes D, Rameh LE (2010) A novel HPLC-based approach makes possible the spatial characterization of cellular PtdIns5P and other phosphoinositides. Biochem J 428(3):375–384 5. Clark J, Anderson KE, Juvin V, Smith TS, Karpe F, Wakelam MJ, Stephens LR, Hawkins PT (2011) Quantification of PtdInsP3 molecular species in cells and tissues by mass spectrometry. Nat Methods 8(3):267–272 6. Whitman M, Downes CP, Keeler M, Keller T, Cantley L (1988) Type I phosphatidylinositol kinase makes a novel inositol phospholipid, phosphatidylinositol-3-phosphate. Nature 332(6165):644–646 7. Devereaux K, Dall’Armi C, Alcazar-Roman A, Ogasawara Y, Zhou X, Wang F, Yamamoto A, De Camilli P, Di Paolo G (2013) Regulation of mammalian autophagy by class II and III PI 3kinases through PI3P synthesis. PLoS One 8(10):e76405 8. Grainger DL, Tavelis C, Ryan AJ, Hinchliffe KA (2012) The emerging role of PtdIns5P: another signalling phosphoinositide takes its place. Biochem Soc Trans 40(1):257–261 9. Zolov SN, Bridges D, Zhang Y, Lee WW, Riehle E, Verma R, Lenk GM, Converso-Baran K, Weide T, Albin RL, Saltiel AR, Meisler MH, Russell MW, Weisman LS (2012) In vivo, Pikfyve generates PI(3,5)P2, which serves as both a signaling lipid and the major precursor for PI5P. Proc Natl Acad Sci U S A 109(43):17472– 17477 10. Funakoshi Y, Hasegawa H, Kanaho Y (2011) Regulation of PIP5K activity by Arf6 and its physiological significance. J Cell Physiol 226(4):888–895 11. Vaccari I, Dina G, Tronchere H, Kaufman E, Chicanne G, Cerri F, Wrabetz L, Payrastre B, Quattrini A, Weisman LS, Meisler MH, Bolino A (2011) Genetic interaction between MTMR2 and FIG4 phospholipid phosphatases involved in Charcot–Marie–Tooth neuropathies. PLoS Genet 7(10):e1002319

12. Chang-Ileto B, Frere SG, Chan RB, Voronov SV, Roux A, Di Paolo G (2011) Synaptojanin 1-mediated PI(4,5)P2 hydrolysis is modulated by membrane curvature and facilitates membrane fission. Dev Cell 20(2):206–218 13. Clarke JH, Emson PC, Irvine RF (2009) Distribution and neuronal expression of phosphatidylinositol phosphate kinase IIgamma in the mouse brain. J Comp Neurol 517(3):296–312 14. Merlot S, Meili R, Pagliarini DJ, Maehama T, Dixon JE, Firtel RA (2003) A PTEN-related 5-phosphatidylinositol phosphatase localized in the Golgi. J Biol Chem 278(41):39866–39873 15. Pagliarini DJ, Worby CA, Dixon JE (2004) A PTEN-like phosphatase with a novel substrate specificity. J Biol Chem 279(37):38590–38596 16. Cain RJ, Ridley AJ (2009) Phosphoinositide 3-kinases in cell migration. Biol Cell 101(1):13–29 17. Sasaki J, Kofuji S, Itoh R, Momiyama T, Takayama K, Murakami H, Chida S, Tsuya Y, Takasuga S, Eguchi S, Asanuma K, Horie Y, Miura K, Davies EM, Mitchell C, Yamazaki M, Hirai H, Takenawa T, Suzuki A, Sasaki T (2010) The PtdIns(3,4)P(2) phosphatase INPP4A is a suppressor of excitotoxic neuronal death. Nature 465(7297):497–501 18. Rahdar M, Inoue T, Meyer T, Zhang J, Vazquez F, Devreotes PN (2009) A phosphorylation-dependent intramolecular interaction regulates the membrane association and activity of the tumor suppressor PTEN. Proc Natl Acad Sci U S A 106(2):480–485 19. Aoki K, Nakamura T, Inoue T, Meyer T, Matsuda M (2007) An essential role for the SHIP2-dependent negative feedback loop in neuritogenesis of nerve growth factor-stimulated PC12 cells. J Cell Biol 177(5):817–827 20. Kisseleva MV, Wilson MP, Majerus PW (2000) The isolation and characterization of a cDNA encoding phospholipid-specific inositol polyphosphate 5-phosphatase. J Biol Chem 275(26):20110–20116 21. Berger P, Bonneick S, Willi S, Wymann M, Suter U (2002) Loss of phosphatase activity in myotubularin-related protein 2 is associated with Charcot–Marie–Tooth disease type 4B1. Hum Mol Genet 11(13):1569–1579 22. Chow CY, Zhang Y, Dowling JJ, Jin N, Adamska M, Shiga K, Szigeti K, Shy ME, Li J, Zhang X, Lupski JR, Weisman LS, Meisler MH (2007) Mutation of FIG4 causes neurodegeneration in the pale tremor mouse and patients with CMT4J. Nature 448(7149):68–72 23. Hokin LE, Hokin MR (1955) Effects of acetylcholine on the turnover of phosphoryl units in individual phospholipids of pancreas slices and brain cortex slices. Biochim Biophys Acta 18(1):102–110 24. Berridge MJ, Irvine RF (1984) Inositol trisphosphate, a novel second messenger in cellular signal transduction. Nature 312(5992):315–321 25. Prakriya M (2013) Store-operated Orai channels: structure and function. Curr Top Membr 71:1–32 26. Berridge MJ (1993) Inositol trisphosphate and calcium signalling. Nature 361(6410):315–325 27. Takai Y, Kishimoto A, Kikkawa U, Mori T, Nishizuka Y (1979) Unsaturated diacylglycerol as a possible messenger for the activation of calcium-activated, phospholipiddependent protein kinase system. Biochem Biophys Res Commun 91(4):1218–1224 28. Kishimoto A, Takai Y, Mori T, Kikkawa U, Nishizuka Y (1980) Activation of calcium and phospholipid-dependent protein kinase by diacylglycerol, its possible relation to phosphatidylinositol turnover. J Biol Chem 255(6):2273–2276 29. Kikkawa U, Takai Y, Tanaka Y, Miyake R, Nishizuka Y (1983) Protein kinase C as a possible receptor protein of tumor-promoting phorbol esters. J Biol Chem 258(19):11442–11445 30. Eberhard DA, Cooper CL, Low MG, Holz RW (1990) Evidence that the inositol phospholipids are necessary for exocytosis. Loss of inositol phospholipids and inhibition of secretion in permeabilized cells caused by a bacterial phospholipase C and removal of ATP. Biochem J 268(1):15–25

Mol Neurobiol 31. Hay JC, Fisette PL, Jenkins GH, Fukami K, Takenawa T, Anderson RA, Martin TF (1995) ATP-dependent inositide phosphorylation required for Ca(2+)-activated secretion. Nature 374(6518):173–177 32. Holz RW, Hlubek MD, Sorensen SD, Fisher SK, Balla T, Ozaki S, Prestwich GD, Stuenkel EL, Bittner MA (2000) A pleckstrin homology domain specific for phosphatidylinositol 4, 5-bisphosphate (PtdIns-4,5-P2) and fused to green fluorescent protein identifies plasma membrane PtdIns-4,5-P2 as being important in exocytosis. J Biol Chem 275(23):17878– 17885 33. Harlan JE, Hajduk PJ, Yoon HS, Fesik SW (1994) Pleckstrin homology domains bind to phosphatidylinositol-4,5-bisphosphate. Nature 371(6493):168–170 34. Traynor-Kaplan AE, Harris AL, Thompson BL, Taylor P, Sklar LA (1988) An inositol tetrakisphosphate-containing phospholipid in activated neutrophils. Nature 334(6180):353–356 35. Auger KR, Serunian LA, Soltoff SP, Libby P, Cantley LC (1989) PDGF-dependent tyrosine phosphorylation stimulates production of novel polyphosphoinositides in intact cells. Cell 57(1):167–175 36. Rameh LE, Tolias KF, Duckworth BC, Cantley LC (1997) A new pathway for synthesis of phosphatidylinositol-4,5-bisphosphate. Nature 390(6656):192–196 37. Whiteford CC, Brearley CA, Ulug ET (1997) Phosphatidylinositol 3,5-bisphosphate defines a novel PI 3-kinase pathway in resting mouse fibroblasts. Biochem J 323(Pt 3):597–601 38. Michell RH, Heath VL, Lemmon MA, Dove SK (2006) Phosphatidylinositol 3,5-bisphosphate: metabolism and cellular functions. Trends Biochem Sci 31(1):52–63 39. Stephens LR, Jackson TR, Hawkins PT (1993) Agonist-stimulated synthesis of phosphatidylinositol(3,4,5)-trisphosphate: a new intracellular signalling system? Biochim Biophys Acta 1179(1):27–75 40. Pendaries C, Tronchere H, Racaud-Sultan C, Gaits-Iacovoni F, Coronas S, Manenti S, Gratacap MP, Plantavid M, Payrastre B (2005) Emerging roles of phosphatidylinositol monophosphates in cellular signaling and trafficking. Adv Enzyme Regul 45:201–214 41. Cremona O, Di Paolo G, Wenk MR, Luthi A, Kim WT, Takei K, Daniell L, Nemoto Y, Shears SB, Flavell RA, McCormick DA, De Camilli P (1999) Essential role of phosphoinositide metabolism in synaptic vesicle recycling. Cell 99(2):179–188 42. Kwon CH, Luikart BW, Powell CM, Zhou J, Matheny SA, Zhang W, Li Y, Baker SJ, Parada LF (2006) PTEN regulates neuronal arborization and social interaction in mice. Neuron 50(3):377–388 43. Antonsson B (1997) Phosphatidylinositol synthase from mammalian tissues. Biochim Biophys Acta 1348(1–2):179–186 44. D’Angelo G, Uemura T, Chuang CC, Polishchuk E, Santoro M, Ohvo-Rekila H, Sato T, Di Tullio G, Varriale A, D’Auria S, Daniele T, Capuani F, Johannes L, Mattjus P, Monti M, Pucci P, Williams RL, Burke JE, Platt FM, Harada A, De Matteis MA (2013) Vesicular and non-vesicular transport feed distinct glycosylation pathways in the Golgi. Nature 501(7465):116–120 45. Guo J, Wenk MR, Pellegrini L, Onofri F, Benfenati F, De Camilli P (2003) Phosphatidylinositol 4-kinase type IIalpha is responsible for the phosphatidylinositol 4-kinase activity associated with synaptic vesicles. Proc Natl Acad Sci U S A 100(7):3995–4000 46. Salazar G, Craige B, Wainer BH, Guo J, De Camilli P, Faundez V (2005) Phosphatidylinositol-4-kinase type II alpha is a component of adaptor protein-3-derived vesicles. Mol Biol Cell 16(8):3692–3704 47. Xu Z, Huang G, Kandror KV (2006) Phosphatidylinositol 4-kinase type IIalpha is targeted specifically to cellugyrin-positive glucose transporter 4 vesicles. Mol Endocrinol 20(11):2890–2897 48. Balla A, Tuymetova G, Barshishat M, Geiszt M, Balla T (2002) Characterization of type II phosphatidylinositol 4-kinase isoforms reveals association of the enzymes with endosomal vesicular compartments. J Biol Chem 277(22):20041–20050 49. Wei YJ, Sun HQ, Yamamoto M, Wlodarski P, Kunii K, Martinez M, Barylko B, Albanesi JP, Yin HL (2002) Type II phosphatidylinositol

50.

51.

52.

53. 54.

55.

56.

57.

58. 59.

60.

61.

62.

63.

64. 65.

66.

67.

4-kinase beta is a cytosolic and peripheral membrane protein that is recruited to the plasma membrane and activated by Rac-GTP. J Biol Chem 277(48):46586–46593 Godi A, Pertile P, Meyers R, Marra P, Di Tullio G, Iurisci C, Luini A, Corda D, De Matteis MA (1999) ARF mediates recruitment of PtdIns-4-OH kinase-beta and stimulates synthesis of PtdIns(4,5)P2 on the Golgi complex. Nat Cell Biol 1(5):280–287 Wong K, Meyers d R, Cantley LC (1997) Subcellular locations of phosphatidylinositol 4-kinase isoforms. J Biol Chem 272(20): 13236–13241 Balla A, Tuymetova G, Tsiomenko A, Varnai P, Balla T (2005) A plasma membrane pool of phosphatidylinositol 4-phosphate is generated by phosphatidylinositol 4-kinase type-III alpha: studies with the PH domains of the oxysterol binding protein and FAPP1. Mol Biol Cell 16(3):1282–1295 Di Paolo G, De Camilli P (2006) Phosphoinositides in cell regulation and membrane dynamics. Nature 443(7112):651–657 Hirst J, Motley A, Harasaki K, Peak Chew SY, Robinson MS (2003) EpsinR: an ENTH domain-containing protein that interacts with AP-1. Mol Biol Cell 14(2):625–641 Wang YJ, Wang J, Sun HQ, Martinez M, Sun YX, Macia E, Kirchhausen T, Albanesi JP, Roth MG, Yin HL (2003) Phosphatidylinositol 4 phosphate regulates targeting of clathrin adaptor AP-1 complexes to the Golgi. Cell 114(3):299–310 Hanada K, Kumagai K, Yasuda S, Miura Y, Kawano M, Fukasawa M, Nishijima M (2003) Molecular machinery for non-vesicular trafficking of ceramide. Nature 426(6968):803–809 Mesmin B, Bigay J, Moser von Filseck J, Lacas-Gervais S, Drin G, Antonny B (2013) A four-step cycle driven by PI(4)P hydrolysis directs sterol/PI(4)P exchange by the ER-Golgi tether OSBP. Cell 155(4):830–843 Frere SG, Chang-Ileto B, Di Paolo G (2012) Role of phosphoinositides at the neuronal synapse. Subcell Biochem 59:131–175 Kawano T, Indo Y, Nakazato H, Shimadzu M, Matsuda I (1998) Oculocerebrorenal syndrome of Lowe: three mutations in the OCRL1 gene derived from three patients with different phenotypes. Am J Med Genet 77(5):348–355 Voronov SV, Frere SG, Giovedi S, Pollina EA, Borel C, Zhang H, Schmidt C, Akeson EC, Wenk MR, Cimasoni L, Arancio O, Davisson MT, Antonarakis SE, Gardiner K, De Camilli P, Di Paolo G (2008) Synaptojanin 1-linked phosphoinositide dyshomeostasis and cognitive deficits in mouse models of Down’s syndrome. Proc Natl Acad Sci U S A 105(27):9415–9420 Wenk MR, Pellegrini L, Klenchin VA, Di Paolo G, Chang S, Daniell L, Arioka M, Martin TF, De Camilli P (2001) PIP kinase Igamma is the major PI(4,5)P(2) synthesizing enzyme at the synapse. Neuron 32(1):79–88 Hammond GR, Fischer MJ, Anderson KE, Holdich J, Koteci A, Balla T, Irvine RF (2012) PI4P and PI(4,5)P2 are essential but independent lipid determinants of membrane identity. Science 337(6095):727–730 Micheva KD, Holz RW, Smith SJ (2001) Regulation of presynaptic phosphatidylinositol 4,5-biphosphate by neuronal activity. J Cell Biol 154(2):355–368 Dittman J, Ryan TA (2009) Molecular circuitry of endocytosis at nerve terminals. Annu Rev Cell Dev Biol 25:133–160 Selak S, Paternain AV, Fritzler MJ, Lerma J (2006) Human autoantibodies against early endosome antigen-1 enhance excitatory synaptic transmission. Neuroscience 143(4):953–964 Tsuruta F, Green EM, Rousset M, Dolmetsch RE (2009) PIKfyve regulates CaV1.2 degradation and prevents excitotoxic cell death. J Cell Biol 187(2):279–294 Seebohm G, Neumann S, Theiss C, Novkovic T, Hill EV, Tavare JM, Lang F, Hollmann M, Manahan-Vaughan D, Strutz-Seebohm N

Mol Neurobiol (2012) Identification of a novel signaling pathway and its relevance for GluA1 recycling. PLoS One 7(3):e33889 68. Zhang Y, McCartney AJ, Zolov SN, Ferguson CJ, Meisler MH, Sutton MA, Weisman LS (2012) Modulation of synaptic function by VAC14, a protein that regulates the phosphoinositides PI(3, 5)P(2) and PI(5)P. EMBO J 31(16):3442–3456 69. Mizutani R, Nakamura K, Yokoyama S, Sanbe A, Kusakawa S, Miyamoto Y, Torii T, Asahara H, Okado H, Yamauchi J, Tanoue A (2011) Developmental expression of sorting nexin 3 in the mouse central nervous system. Gene Expr Patterns 11(1–2):33–40 70. Ackermann M, Matus A (2003) Activity-induced targeting of profilin and stabilization of dendritic spine morphology. Nat Neurosci 6(11):1194–1200 71. Gu J, Lee CW, Fan Y, Komlos D, Tang X, Sun C, Yu K, Hartzell HC, Chen G, Bamburg JR, Zheng JQ (2010) ADF/cofilin-mediated actin dynamics regulate AMPA receptor trafficking during synaptic plasticity. Nat Neurosci 13(10):1208–1215 72. Wegner AM, Nebhan CA, Hu L, Majumdar D, Meier KM, Weaver AM, Webb DJ (2008) N-wasp and the arp2/3 complex are critical regulators of actin in the development of dendritic spines and synapses. J Biol Chem 283(23):15912–15920 73. Furukawa K, Fu W, Li Y, Witke W, Kwiatkowski DJ, Mattson MP (1997) The actin-severing protein gelsolin modulates calcium channel and NMDA receptor activities and vulnerability to excitotoxicity in hippocampal neurons. J Neurosci 17(21):8178–8186 74. Di Paolo G, Pellegrini L, Letinic K, Cestra G, Zoncu R, Voronov S, Chang S, Guo J, Wenk MR, De Camilli P (2002) Recruitment and regulation of phosphatidylinositol phosphate kinase type 1 gamma by the FERM domain of talin. Nature 420(6911):85–89 75. Kim HS, Bae CD, Park J (2010) Glutamate receptor-mediated phosphorylation of ezrin/radixin/moesin proteins is implicated in filopodial protrusion of primary cultured hippocampal neuronal cells. J Neurochem 113(6):1565–1576 76. Antoine-Bertrand J, Ghogha A, Luangrath V, Bedford FK, Lamarche-Vane N (2011) The activation of ezrin-radixin-moesin proteins is regulated by netrin-1 through Src kinase and RhoA/Rho kinase activities and mediates netrin-1-induced axon outgrowth. Mol Biol Cell 22(19):3734–3746 77. Marsick BM, San Miguel-Ruiz JE, Letourneau PC (2012) Activation of ezrin/radixin/moesin mediates attractive growth cone guidance through regulation of growth cone actin and adhesion receptors. J Neurosci 32(1):282–296 78. Fox JW, Lamperti ED, Eksioglu YZ, Hong SE, Feng Y, Graham DA, Scheffer IE, Dobyns WB, Hirsch BA, Radtke RA, Berkovic SF, Huttenlocher PR, Walsh CA (1998) Mutations in filamin 1 prevent migration of cerebral cortical neurons in human periventricular heterotopia. Neuron 21(6): 1315–1325 79. Hoe HS, Lee JY, Pak DT (2009) Combinatorial morphogenesis of dendritic spines and filopodia by SPAR and alpha-actinin2. Biochem Biophys Res Commun 384(1):55–60 80. Michailidis IE, Helton TD, Petrou VI, Mirshahi T, Ehlers MD, Logothetis DE (2007) Phosphatidylinositol-4,5-bisphosphate regulates NMDA receptor activity through alpha-actinin. J Neurosci 27(20):5523–5532 81. Matus A (2005) MARCKS for maintenance in dendritic spines. Neuron 48(1):4–5 82. Hussain RJ, Stumpo DJ, Blackshear PJ, Lenox RH, Abel T, McNamara RK (2006) Myristoylated alanine rich C kinase substrate (MARCKS) heterozygous mutant mice exhibit deficits in hippocampal mossy fiber-CA3 long-term potentiation. Hippocampus 16(5):495–503 83. Chen YK, Hsueh YP (2012) Cortactin-binding protein 2 modulates the mobility of cortactin and regulates dendritic spine formation and maintenance. J Neurosci 32(3):1043–1055

84. Yamada H, Abe T, Satoh A, Okazaki N, Tago S, Kobayashi K, Yoshida Y, Oda Y, Watanabe M, Tomizawa K, Matsui H, Takei K (2013) Stabilization of actin bundles by a dynamin 1/cortactin ring complex is necessary for growth cone filopodia. J Neurosci 33(10): 4514–4526 85. De Vos KJ, Sable J, Miller KE, Sheetz MP (2003) Expression of phosphatidylinositol (4,5) bisphosphate-specific pleckstrin homology domains alters direction but not the level of axonal transport of mitochondria. Mol Biol Cell 14(9):3636–3649 86. Gao Y, Perkins EM, Clarkson YL, Tobia S, Lyndon AR, Jackson M, Rothstein JD (2011) beta-III spectrin is critical for development of purkinje cell dendritic tree and spine morphogenesis. J Neurosci 31(46):16581–16590 87. Medina DL, Sciarretta C, Calella AM, Von Bohlen Und Halbach O, Unsicker K, Minichiello L (2004) TrkB regulates neocortex formation through the Shc/PLCgamma-mediated control of neuronal migration. EMBO J 23(19):3803–3814 88. Jahn R, Fasshauer D (2012) Molecular machines governing exocytosis of synaptic vesicles. Nature 490(7419):201–207 89. Yao J, Gaffaney JD, Kwon SE, Chapman ER (2011) Doc2 is a Ca2+ sensor required for asynchronous neurotransmitter release. Cell 147(3):666–677 90. Augustin I, Rosenmund C, Sudhof TC, Brose N (1999) Munc13-1 is essential for fusion competence of glutamatergic synaptic vesicles. Nature 400(6743):457–461 91. Chung SH, Takai Y, Holz RW (1995) Evidence that the Rab3abinding protein, rabphilin3a, enhances regulated secretion. Studies in adrenal chromaffin cells. J Biol Chem 270(28):16714–16718 92. Jockusch WJ, Speidel D, Sigler A, Sorensen JB, Varoqueaux F, Rhee JS, Brose N (2007) CAPS-1 and CAPS-2 are essential synaptic vesicle priming proteins. Cell 131(4):796–808 93. Okamoto M, Sudhof TC (1997) Mints, Munc18-interacting proteins in synaptic vesicle exocytosis. J Biol Chem 272(50):31459–31464 94. Wang X, Hu B, Zieba A, Neumann NG, Kasper-Sonnenberg M, Honsbein A, Hultqvist G, Conze T, Witt W, Limbach C, Geitmann M, Danielson H, Kolarow R, Niemann G, Lessmann V, Kilimann MW (2009) A protein interaction node at the neurotransmitter release site: domains of Aczonin/Piccolo, Bassoon, CAST, and rim converge on the N-terminal domain of Munc13-1. J Neurosci 29(40):12584–12596 95. Schoch S, Castillo PE, Jo T, Mukherjee K, Geppert M, Wang Y, Schmitz F, Malenka RC, Sudhof TC (2002) RIM1alpha forms a protein scaffold for regulating neurotransmitter release at the active zone. Nature 415(6869):321–326 96. Koushika SP, Richmond JE, Hadwiger G, Weimer RM, Jorgensen EM, Nonet ML (2001) A post-docking role for active zone protein Rim. Nat Neurosci 4(10):997–1005 97. van den Bogaart G, Meyenberg K, Risselada HJ, Amin H, Willig KI, Hubrich BE, Dier M, Hell SW, Grubmuller H, Diederichsen U, Jahn R (2011) Membrane protein sequestering by ionic protein– lipid interactions. Nature 479(7374):552–555 98. Jakobsson J, Gad H, Andersson F, Low P, Shupliakov O, Brodin L (2008) Role of epsin 1 in synaptic vesicle endocytosis. Proc Natl Acad Sci U S A 105(17):6445–6450 99. Krauss M, Kinuta M, Wenk MR, De Camilli P, Takei K, Haucke V (2003) ARF6 stimulates clathrin/AP-2 recruitment to synaptic membranes by activating phosphatidylinositol phosphate kinase type Igamma. J Cell Biol 162(1):113–124 100. Ford MG, Pearse BM, Higgins MK, Vallis Y, Owen DJ, Gibson A, Hopkins CR, Evans PR, McMahon HT (2001) Simultaneous binding of PtdIns(4,5)P2 and clathrin by AP180 in the nucleation of clathrin lattices on membranes. Science 291(5506):1051–1055 101. Jackson LP, Kelly BT, McCoy AJ, Gaffry T, James LC, Collins BM, Honing S, Evans PR, Owen DJ (2010) A large-scale conformational change couples membrane recruitment to cargo binding in the AP2 clathrin adaptor complex. Cell 141(7):1220–1229

Mol Neurobiol 102. Takei K, Mundigl O, Daniell L, De Camilli P (1996) The synaptic vesicle cycle: a single vesicle budding step involving clathrin and dynamin. J Cell Biol 133(6):1237–1250 103. Zheng J, Cahill SM, Lemmon MA, Fushman D, Schlessinger J, Cowburn D (1996) Identification of the binding site for acidic phospholipids on the pH domain of dynamin: implications for stimulation of GTPase activity. J Mol Biol 255(1):14–21 104. Lu J, Helton TD, Blanpied TA, Racz B, Newpher TM, Weinberg RJ, Ehlers MD (2007) Postsynaptic positioning of endocytic zones and AMPA receptor cycling by physical coupling of dynamin-3 to Homer. Neuron 55(6):874–889 105. Gonzalez-Jamett AM, Haro-Acuna V, Momboisse F, Caviedes P, Bevilacqua JA, Cardenas AM (2014) Dynamin-2 in nervous system disorders. J Neurochem 128(2):210–223 106. Kanaho Y, Funakoshi Y, Hasegawa H (2009) Phospholipase D signalling and its involvement in neurite outgrowth. Biochim Biophys Acta 1791(9):898–904 107. Lee HW, Choi J, Shin H, Kim K, Yang J, Na M, Choi SY, Kang GB, Eom SH, Kim H, Kim E (2008) Preso, a novel PSD-95-interacting FERM and PDZ domain protein that regulates dendritic spine morphogenesis. J Neurosci 28(53): 14546–14556 108. Mo J, Lee D, Hong S, Han S, Yeo H, Sun W, Choi S, Kim H, Lee HW (2012) Preso regulation of dendritic outgrowth through PI(4, 5)P2-dependent PDZ interaction with betaPix. Eur J Neurosci 36(1): 1960–1970 109. Zhang H, Craciun LC, Mirshahi T, Rohacs T, Lopes CM, Jin T, Logothetis DE (2003) PIP(2) activates KCNQ channels, and its hydrolysis underlies receptor-mediated inhibition of M currents. Neuron 37(6):963–975 110. van Gisbergen PA, Li M, Wu SZ, Bezanilla M (2012) Class II formin targeting to the cell cortex by binding PI(3,5)P(2) is essential for polarized growth. J Cell Biol 198(2):235–250 111. Hernandez-Deviez DJ, Casanova JE, Wilson JM (2002) Regulation of dendritic development by the ARF exchange factor ARNO. Nat Neurosci 5(7):623–624 112. Miyamoto Y, Chen L, Sato M, Sokabe M, Nabeshima T, Pawson T, Sakai R, Mori N (2005) Hippocampal synaptic modulation by the phosphotyrosine adapter protein ShcC/N-Shc via interaction with the NMDA receptor. J Neurosci 25(7):1826–1835 113. Tolias KF, Bikoff JB, Kane CG, Tolias CS, Hu L, Greenberg ME (2007) The Rac1 guanine nucleotide exchange factor Tiam1 mediates EphB receptor-dependent dendritic spine development. Proc Natl Acad Sci U S A 104(17):7265–7270 114. Arai JA, Li S, Feig LA (2009) Sos2 is dispensable for NMDAinduced Erk activation and LTP induction. Neurosci Lett 455(1):22– 25 115. Jaworski J, Spangler S, Seeburg DP, Hoogenraad CC, Sheng M (2005) Control of dendritic arborization by the phosphoinositide-3′kinase-Akt-mammalian target of rapamycin pathway. J Neurosci 25(49):11300–11312 116. Pantakani DV, Czyzewska MM, Sikorska A, Bodda C, Mannan AU (2011) Oligomerization of ZFYVE27 (Protrudin) is necessary to promote neurite extension. PLoS One 6(12):e29584 117. Gil JE, Kim E, Kim IS, Ku B, Park WS, Oh BH, Ryu SH, Cho W, Heo WD (2012) Phosphoinositides differentially regulate protrudin localization through the FYVE domain. J Biol Chem 287(49): 41268–41276 118. Vicinanza M, D’Angelo G, Di Campli A, De Matteis MA (2008) Function and dysfunction of the PI system in membrane trafficking. EMBO J 27(19):2457–2470 119. van Diepen MT, Parsons M, Downes CP, Leslie NR, Hindges R, Eickholt BJ (2009) MyosinV controls PTEN function and neuronal cell size. Nat Cell Biol 11(10):1191–1196 120. Ketschek A, Gallo G (2010) Nerve growth factor induces axonal filopodia through localized microdomains of phosphoinositide 3-

kinase activity that drive the formation of cytoskeletal precursors to filopodia. J Neurosci 30(36):12185–12197 121. Sanna PP, Cammalleri M, Berton F, Simpson C, Lutjens R, Bloom FE, Francesconi W (2002) Phosphatidylinositol 3-kinase is required for the expression but not for the induction or the maintenance of long-term potentiation in the hippocampal CA1 region. J Neurosci 22(9):3359–3365 122. Arendt KL, Royo M, Fernandez-Monreal M, Knafo S, Petrok CN, Martens JR, Esteban JA (2010) PIP3 controls synaptic function by maintaining AMPA receptor clustering at the postsynaptic membrane. Nat Neurosci 13(1):36–44 123. Khuong TM, Habets RL, Kuenen S, Witkowska A, Kasprowicz J, Swerts J, Jahn R, van den Bogaart G, Verstreken P (2013) Synaptic PI(3,4,5)P3 is required for Syntaxin1A clustering and neurotransmitter release. Neuron 77(6):1097–1108 124. Man HY, Wang Q, Lu WY, Ju W, Ahmadian G, Liu L, D’Souza S, Wong TP, Taghibiglou C, Lu J, Becker LE, Pei L, Liu F, Wymann MP, MacDonald JF, Wang YT (2003) Activation of PI3-kinase is required for AMPA receptor insertion during LTP of mEPSCs in cultured hippocampal neurons. Neuron 38(4):611–624 125. van Diepen MT, Eickholt BJ (2008) Function of PTEN during the formation and maintenance of neuronal circuits in the brain. Dev Neurosci 30(1–3):59–64 126. Ueda Y, Hayashi Y (2013) PIP(3) regulates spinule formation in dendritic spines during structural long-term potentiation. J Neurosci 33(27):11040–11047 127. Bolino A, Muglia M, Conforti FL, LeGuern E, Salih MA, Georgiou DM, Christodoulou K, Hausmanowa-Petrusewicz I, Mandich P, Schenone A, Gambardella A, Bono F, Quattrone A, Devoto M, Monaco AP (2000) Charcot–Marie–Tooth type 4B is caused by mutations in the gene encoding myotubularin-related protein-2. Nat Genet 25(1):17–19 128. Cooke FT (2002) Phosphatidylinositol 3,5-bisphosphate: metabolism and function. Arch Biochem Biophys 407(2):143–151 129. Clarke JH, Irvine RF (2013) Evolutionarily conserved structural changes in phosphatidylinositol 5-phosphate 4-kinase (PI5P4K) isoforms are responsible for differences in enzyme activity and localization. Biochem J 454(1):49–57 130. Guittard G, Gerard A, Dupuis-Coronas S, Tronchere H, Mortier E, Favre C, Olive D, Zimmermann P, Payrastre B, Nunes JA (2009) Cutting edge: Dok-1 and Dok-2 adaptor molecules are regulated by phosphatidylinositol 5-phosphate production in T cells. J Immunol 182(7):3974–3978 131. Guittard G, Mortier E, Tronchere H, Firaguay G, Gerard A, Zimmermann P, Payrastre B, Nunes JA (2010) Evidence for a positive role of PtdIns5P in T-cell signal transduction pathways. FEBS Lett 584(11):2455–2460 132. Gozani O, Karuman P, Jones DR, Ivanov D, Cha J, Lugovskoy AA, Baird CL, Zhu H, Field SJ, Lessnick SL, Villasenor J, Mehrotra B, Chen J, Rao VR, Brugge JS, Ferguson CG, Payrastre B, Myszka DG, Cantley LC, Wagner G, Divecha N, Prestwich GD, Yuan J (2003) The PHD finger of the chromatin-associated protein ING2 functions as a nuclear phosphoinositide receptor. Cell 114(1):99– 111 133. Lee HW, Kim Y, Han K, Kim H, Kim E (2010) The phosphoinositide 3-phosphatase MTMR2 interacts with PSD-95 and maintains excitatory synapses by modulating endosomal traffic. J Neurosci 30(16):5508–5518 134. Nystuen A, Legare ME, Shultz LD, Frankel WN (2001) A null mutation in inositol polyphosphate 4-phosphatase type I causes selective neuronal loss in weeble mutant mice. Neuron 32(2):203– 212 135. Helgason CD, Damen JE, Rosten P, Grewal R, Sorensen P, Chappel SM, Borowski A, Jirik F, Krystal G, Humphries RK (1998) Targeted disruption of SHIP leads to hemopoietic perturbations,

Mol Neurobiol

136.

137.

138.

139. 140.

141.

142.

143.

144.

145.

146.

147.

148.

149.

150.

151.

lung pathology, and a shortened life span. Genes Dev 12(11):1610– 1620 Nishioka T, Aoki K, Hikake K, Yoshizaki H, Kiyokawa E, Matsuda M (2008) Rapid turnover rate of phosphoinositides at the front of migrating MDCK cells. Mol Biol Cell 19(10): 4213–4223 Mishra A, Eathiraj S, Corvera S, Lambright DG (2010) Structural basis for Rab GTPase recognition and endosome tethering by the C2H2 zinc finger of Early Endosomal Autoantigen 1 (EEA1). Proc Natl Acad Sci U S A 107(24):10866–10871 Gaidarov I, Smith ME, Domin J, Keen JH (2001) The class II phosphoinositide 3-kinase C2alpha is activated by clathrin and regulates clathrin-mediated membrane trafficking. Mol Cell 7(2): 443–449 Vergne I, Deretic V (2010) The role of PI3P phosphatases in the regulation of autophagy. FEBS Lett 584(7):1313–1318 Vanhaesebroeck B, Guillermet-Guibert J, Graupera M, Bilanges B (2010) The emerging mechanisms of isoform-specific PI3K signalling. Nat Rev Mol Cell Biol 11(5):329–341 Schu PV, Takegawa K, Fry MJ, Stack JH, Waterfield MD, Emr SD (1993) Phosphatidylinositol 3-kinase encoded by yeast VPS34 gene essential for protein sorting. Science 260(5104):88–91 Wang L, Budolfson K, Wang F (2011) Pik3c3 deletion in pyramidal neurons results in loss of synapses, extensive gliosis and progressive neurodegeneration. Neuroscience 172:427–442 Sbrissa D, Ikonomov OC, Fu Z, Ijuin T, Gruenberg J, Takenawa T, Shisheva A (2007) Core protein machinery for mammalian phosphatidylinositol 3,5-bisphosphate synthesis and turnover that regulates the progression of endosomal transport. Novel Sac phosphatase joins the ArPIKfyve-PIKfyve complex. J Biol Chem 282(33): 23878–23891 Ikonomov OC, Sbrissa D, Fenner H, Shisheva A (2009) PIKfyveArPIKfyve-Sac3 core complex: contact sites and their consequence for Sac3 phosphatase activity and endocytic membrane homeostasis. J Biol Chem 284(51):35794–35806 Nasuhoglu C, Feng S, Mao J, Yamamoto M, Yin HL, Earnest S, Barylko B, Albanesi JP, Hilgemann DW (2002) Nonradioactive analysis of phosphatidylinositides and other anionic phospholipids by anion-exchange high-performance liquid chromatography with suppressed conductivity detection. Anal Biochem 301(2):243–254 Milne SB, Ivanova PT, DeCamp D, Hsueh RC, Brown HA (2005) A targeted mass spectrometric analysis of phosphatidylinositol phosphate species. J Lipid Res 46(8):1796–1802 Wenk MR, Lucast L, Di Paolo G, Romanelli AJ, Suchy SF, Nussbaum RL, Cline GW, Shulman GI, McMurray W, De Camilli P (2003) Phosphoinositide profiling in complex lipid mixtures using electrospray ionization mass spectrometry. Nat Biotechnol 21(7):813–817 Ogiso H, Nakamura K, Yatomi Y, Shimizu T, Taguchi R (2010) Liquid chromatography/mass spectrometry analysis revealing preferential occurrence of non-arachidonate-containing phosphatidylinositol bisphosphate species in nuclei and changes in their levels during cell cycle. Rapid Commun Mass Spectrom 24(4):436– 442 Matsuura D, Taguchi K, Yagisawa H, Maekawa S (2007) Lipid components in the detergent-resistant membrane microdomain (DRM) obtained from the synaptic plasma membrane of rat brain. Neurosci Lett 423(2):158–161 Schiavo G, Gu QM, Prestwich GD, Sollner TH, Rothman JE (1996) Calcium-dependent switching of the specificity of phosphoinositide binding to synaptotagmin. Proc Natl Acad Sci U S A 93(23):13327– 13332 Varnai P, Rother KI, Balla T (1999) Phosphatidylinositol 3-kinasedependent membrane association of the Bruton’s tyrosine kinase pleckstrin homology domain visualized in single living cells. J Biol Chem 274(16):10983–10989

152. Venkateswarlu K, Gunn-Moore F, Oatey PB, Tavare JM, Cullen PJ (1998) Nerve growth factor- and epidermal growth factorstimulated translocation of the ADP-ribosylation factor-exchange factor GRP1 to the plasma membrane of PC12 cells requires activation of phosphatidylinositol 3-kinase and the GRP1 pleckstrin homology domain. Biochem J 335(Pt 1):139–146 153. Horne EA, Dell’Acqua ML (2007) Phospholipase C is required for changes in postsynaptic structure and function associated with NMDA receptor-dependent long-term depression. J Neurosci 27(13):3523–3534 154. Sato M, Ueda Y, Takagi T, Umezawa Y (2003) Production of PtdInsP3 at endomembranes is triggered by receptor endocytosis. Nat Cell Biol 5(11):1016–1022 155. Sato M, Ueda Y, Umezawa Y (2006) Imaging diacylglycerol dynamics at organelle membranes. Nat Methods 3(10):797–799 156. Ueda Y, Kwok S, Hayashi Y (2013) Application of FRET probes in the analysis of neuronal plasticity. Front Neural Circ 7:163 157. Bourne JN, Harris KM (2012) Nanoscale analysis of structural synaptic plasticity. Curr Opin Neurobiol 22(3):372–382 158. Spacek J, Harris KM (1997) Three-dimensional organization of smooth endoplasmic reticulum in hippocampal CA1 dendrites and dendritic spines of the immature and mature rat. J Neurosci 17(1): 190–203 159. Park M, Salgado JM, Ostroff L, Helton TD, Robinson CG, Harris KM, Ehlers MD (2006) Plasticity-induced growth of dendritic spines by exocytic trafficking from recycling endosomes. Neuron 52(5):817–830 160. Lynch MA (2004) Long-term potentiation and memory. Physiol Rev 84(1):87–136 161. Liscovitch M, Cantley LC (1995) Signal transduction and membrane traffic: the PITP/phosphoinositide connection. Cell 81(5): 659–662 162. Martens S, Kozlov MM, McMahon HT (2007) How synaptotagmin promotes membrane fusion. Science 316(5828):1205–1208 163. Sudhof TC (2004) The synaptic vesicle cycle. Annu Rev Neurosci 27:509–547 164. Sudhof TC, Rizo J (2011) Synaptic vesicle exocytosis. Cold Spring Harb Perspect Biol 3 (12) 165. Burgalossi A, Jung S, Meyer G, Jockusch WJ, Jahn O, Taschenberger H, O’Connor VM, Nishiki T, Takahashi M, Brose N, Rhee JS (2010) SNARE protein recycling by alphaSNAP and betaSNAP supports synaptic vesicle priming. Neuron 68(3):473– 487 166. Martin TF (2012) Role of PI(4,5)P(2) in vesicle exocytosis and membrane fusion. Subcell Biochem 59:111–130 167. Kasai H, Takahashi N, Tokumaru H (2012) Distinct initial SNARE configurations underlying the diversity of exocytosis. Physiol Rev 92(4):1915–1964 168. Haucke V, Neher E, Sigrist SJ (2011) Protein scaffolds in the coupling of synaptic exocytosis and endocytosis. Nat Rev Neurosci 12(3):127–138 169. Shin OH, Lu J, Rhee JS, Tomchick DR, Pang ZP, Wojcik SM, Camacho-Perez M, Brose N, Machius M, Rizo J, Rosenmund C, Sudhof TC (2010) Munc13 C2B domain is an activity-dependent Ca2+ regulator of synaptic exocytosis. Nat Struct Mol Biol 17(3): 280–288 170. Wang Z, Liu H, Gu Y, Chapman ER (2011) Reconstituted synaptotagmin I mediates vesicle docking, priming, and fusion. J Cell Biol 195(7):1159–1170 171. Kessels HW, Malinow R (2009) Synaptic AMPA receptor plasticity and behavior. Neuron 61(3):340–350 172. Beattie EC, Carroll RC, Yu X, Morishita W, Yasuda H, von Zastrow M, Malenka RC (2000) Regulation of AMPA receptor endocytosis by a signaling mechanism shared with LTD. Nat Neurosci 3(12): 1291–1300

Mol Neurobiol 173. Huganir RL, Nicoll RA (2013) AMPARs and synaptic plasticity: the Last 25 years. Neuron 80(3):704–717 174. Tada T, Sheng M (2006) Molecular mechanisms of dendritic spine morphogenesis. Curr Opin Neurobiol 16(1):95–101 175. Fiala JC, Spacek J, Harris KM (2002) Dendritic spine pathology: cause or consequence of neurological disorders? Brain Res Brain Res Rev 39(1):29–54 176. Toni N, Buchs PA, Nikonenko I, Bron CR, Muller D (1999) LTP promotes formation of multiple spine synapses between a single axon terminal and a dendrite. Nature 402(6760):421–425 177. Richards DA, Mateos JM, Hugel S, de Paola V, Caroni P, Gahwiler BH, McKinney RA (2005) Glutamate induces the rapid formation of spine head protrusions in hippocampal slice cultures. Proc Natl Acad Sci U S A 102(17):6166–6171 178. Matsuzaki M, Honkura N, Ellis-Davies GC, Kasai H (2004) Structural basis of long-term potentiation in single dendritic spines. Nature 429(6993):761–766 179. Jurado S, Benoist M, Lario A, Knafo S, Petrok CN, Esteban JA (2010) PTEN is recruited to the postsynaptic terminal for NMDA receptor-dependent long-term depression. EMBO J 29(16):2827– 2840 180. Dell’Acqua ML, Smith KE, Gorski JA, Horne EA, Gibson ES, Gomez LL (2006) Regulation of neuronal PKA signaling through AKAP targeting dynamics. Eur J Cell Biol 85(7):627–633 181. Unoki T, Matsuda S, Kakegawa W, Van NT, Kohda K, Suzuki A, Funakoshi Y, Hasegawa H, Yuzaki M, Kanaho Y (2012) NMDA receptor-mediated PIP5K activation to produce PI(4,5)P(2) is essential for AMPA receptor endocytosis during LTD. Neuron 73(1): 135–148 182. Okamoto K, Nagai T, Miyawaki A, Hayashi Y (2004) Rapid and persistent modulation of actin dynamics regulates postsynaptic reorganization underlying bidirectional plasticity. Nat Neurosci 7(10): 1104–1112 183. Shirao T, Gonzalez-Billault C (2013) Actin filaments and microtubules in dendritic spines. J Neurochem 126(2):155–164 184. Ueno T, Falkenburger BH, Pohlmeyer C, Inoue T (2011) Triggering actin comets versus membrane ruffles: distinctive effects of phosphoinositides on actin reorganization. Sci Sign 4 (203):ra87 185. Lambrechts A, Jonckheere V, Dewitte D, Vandekerckhove J, Ampe C (2002) Mutational analysis of human profilin I reveals a second PI(4,5)-P2 binding site neighbouring the poly(L-proline) binding site. BMC Biochem 3:12 186. Gorbatyuk VY, Nosworthy NJ, Robson SA, Bains NP, Maciejewski MW, Dos Remedios CG, King GF (2006) Mapping the phosphoinositide-binding site on chick cofilin explains how PIP2 regulates the cofilin-actin interaction. Mol Cell 24(4):511–522 187. Lin KM, Wenegieme E, Lu PJ, Chen CS, Yin HL (1997) Gelsolin binding to phosphatidylinositol 4,5-bisphosphate is modulated by calcium and pH. J Biol Chem 272(33):20443–20450 188. Murakoshi H, Yasuda R (2012) Postsynaptic signaling during plasticity of dendritic spines. Trends Neurosci 35(2):135–143 189. Lee SJ, Escobedo-Lozoya Y, Szatmari EM, Yasuda R (2009) Activation of CaMKII in single dendritic spines during long-term potentiation. Nature 458(7236):299–304 190. Murakoshi H, Wang H, Yasuda R (2011) Local, persistent activation of Rho GTPases during plasticity of single dendritic spines. Nature 472(7341):100–104 191. Tao-Cheng JH, Dosemeci A, Gallant PE, Miller S, Galbraith JA, Winters CA, Azzam R, Reese TS (2009) Rapid turnover of spinules at synaptic terminals. Neuroscience 160(1):42–50

192. Spacek J, Harris KM (2004) Trans-endocytosis via spinules in adult rat hippocampus. J Neurosci 24(17):4233–4241 193. Toomre D, Bewersdorf J (2010) A new wave of cellular imaging. Annu Rev Cell Dev Biol 26:285–314 194. Frost NA, Shroff H, Kong H, Betzig E, Blanpied TA (2010) Singlemolecule discrimination of discrete perisynaptic and distributed sites of actin filament assembly within dendritic spines. Neuron 67(1):86–99 195. Izeddin I, Specht CG, Lelek M, Darzacq X, Triller A, Zimmer C, Dahan M (2011) Super-resolution dynamic imaging of dendritic spines using a low-affinity photoconvertible actin probe. PLoS One 6(1):e15611 196. Urban NT, Willig KI, Hell SW, Nagerl UV (2011) STED nanoscopy of actin dynamics in synapses deep inside living brain slices. Biophys J 101(5):1277–1284 197. Bauer M, Alda M, Priller J, Young LT (2003) Implications of the neuroprotective effects of lithium for the treatment of bipolar and neurodegenerative disorders. Pharmacopsychiatry 36(Suppl 3): S250–S254 198. Geddes JR, Miklowitz DJ (2013) Treatment of bipolar disorder. Lancet 381(9878):1672–1682 199. Salinas PC, Hall AC (1999) Lithium and synaptic plasticity. Bipolar Disord 1(2):87–90 200. McBride KL, Varga EA, Pastore MT, Prior TW, Manickam K, Atkin JF, Herman GE (2010) Confirmation study of PTEN mutations among individuals with autism or developmental delays/mental retardation and macrocephaly. Autism Res 3(3):137–141 201. Halstead JR, Jalink K, Divecha N (2005) An emerging role for PtdIns(4,5)P2-mediated signalling in human disease. Trends Pharmacol Sci 26(12):654–660 202. Berman DE, Dall’Armi C, Voronov SV, McIntire LB, Zhang H, Moore AZ, Staniszewski A, Arancio O, Kim TW, Di Paolo G (2008) Oligomeric amyloid-beta peptide disrupts phosphatidylinositol-4,5-bisphosphate metabolism. Nat Neurosci 11(5):547–554 203. McIntire LB, Berman DE, Myaeng J, Staniszewski A, Arancio O, Di Paolo G, Kim TW (2012) Reduction of synaptojanin 1 ameliorates synaptic and behavioral impairments in a mouse model of Alzheimer’s disease. J Neurosci 32(44):15271–15276 204. Zhu L, Zhong M, Zhao J, Rhee H, Caesar I, Knight EM, Volpicelli-Daley L, Bustos V, Netzer W, Liu L, Lucast L, Ehrlich ME, Robakis NK, Gandy SE, Cai D (2013) Reduction of synaptojanin 1 accelerates Abeta clearance and attenuates cognitive deterioration in an Alzheimer mouse model. J Biol Chem 288(44):32050–32063 205. Cossec JC, Lavaur J, Berman DE, Rivals I, Hoischen A, Stora S, Ripoll C, Mircher C, Grattau Y, Olivomarin JC, de Chaumont F, Lecourtois M, Antonarakis SE, Veltman JA, Delabar JM, Duyckaerts C, Di Paolo G, Potier MC (2012) Trisomy for synaptojanin1 in Down syndrome is functionally linked to the enlargement of early endosomes. Hum Mol Genet 21(14):3156–3172 206. Chang KT, Min KT (2009) Upregulation of three Drosophila homologs of human chromosome 21 genes alters synaptic function: implications for Down syndrome. Proc Natl Acad Sci U S A 106(40):17117–17122 207. Herrera F, Chen Q, Fischer WH, Maher P, Schubert DR (2009) Synaptojanin-1 plays a key role in astrogliogenesis: possible relevance for Down’s syndrome. Cell Death Differ 16(6):910–920 208. Arai Y, Ijuin T, Takenawa T, Becker LE, Takashima S (2002) Excessive expression of synaptojanin in brains with Down syndrome. Brain Dev 24(2):67–72

The role of phosphoinositides in synapse function.

Since the discovery of phosphatidylinositol-3-kinase, scientific interest in the biological functions of phosphoinositides has greatly increased. Curr...
807KB Sizes 5 Downloads 3 Views