Next Article in Journal
Treating Cancers Using Nature’s Medicine: Significance and Challenges
Next Article in Special Issue
Integration and Spatial Organization of Signaling by G Protein-Coupled Receptor Homo- and Heterodimers
Previous Article in Journal
Immunotoxins: From Design to Clinical Application
Previous Article in Special Issue
Quantitative Super-Resolution Imaging for the Analysis of GPCR Oligomerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanoscale Sub-Compartmentalization of the Dendritic Spine Compartment

by
Ana Sofía Vallés
1 and
Francisco J. Barrantes
2,*
1
Instituto de Investigaciones Bioquímicas de Bahía Blanca (UNS-CONICET), Bahía Blanca 8000, Argentina
2
Laboratory of Molecular Neurobiology, Institute of Biomedical Research (BIOMED), UCA-CONICET, Av. Alicia Moreau de Justo 1600, Buenos Aires C1107AFF, Argentina
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(11), 1697; https://doi.org/10.3390/biom11111697
Submission received: 21 October 2021 / Revised: 10 November 2021 / Accepted: 11 November 2021 / Published: 15 November 2021
(This article belongs to the Special Issue Molecular Mechanisms of Compartmentalized GPCR Signaling)

Abstract

:
Compartmentalization of the membrane is essential for cells to perform highly specific tasks and spatially constrained biochemical functions in topographically defined areas. These membrane lateral heterogeneities range from nanoscopic dimensions, often involving only a few molecular constituents, to micron-sized mesoscopic domains resulting from the coalescence of nanodomains. Short-lived domains lasting for a few milliseconds coexist with more stable platforms lasting from minutes to days. This panoply of lateral domains subserves the great variety of demands of cell physiology, particularly high for those implicated in signaling. The dendritic spine, a subcellular structure of neurons at the receiving (postsynaptic) end of central nervous system excitatory synapses, exploits this compartmentalization principle. In its most frequent adult morphology, the mushroom-shaped spine harbors neurotransmitter receptors, enzymes, and scaffolding proteins tightly packed in a volume of a few femtoliters. In addition to constituting a mesoscopic lateral heterogeneity of the dendritic arborization, the dendritic spine postsynaptic membrane is further compartmentalized into spatially delimited nanodomains that execute separate functions in the synapse. This review discusses the functional relevance of compartmentalization and nanodomain organization in synaptic transmission and plasticity and exemplifies the importance of this parcelization in various neurotransmitter signaling systems operating at dendritic spines, using two fast ligand-gated ionotropic receptors, the nicotinic acetylcholine receptor and the glutamatergic receptor, and a second-messenger G-protein coupled receptor, the cannabinoid receptor, as paradigmatic examples.

1. Introduction

The plasma membrane is currently envisaged as a highly heterogeneous and compartmentalized two-dimensional lattice. Convergent biochemical and biophysical tools have provided experimental support to this notion (see reviews in [1,2,3]). The finding of coexisting lipid domains with distinct physicochemical properties (liquid-ordered and liquid-disordered phases, respectively) observed in biomimetic synthetic lipid membranes inspired the “lipid raft” hypothesis, which purported that similar co-occurrence of phases also occurred in natural cell membranes, whereby rafts corresponded to cholesterol- and sphingolipid-rich liquid-ordered domains [4,5]. The functional implications of the hypothesis were vast: the early proposal that these lipid domains served as signaling platforms was an appealing idea [6] that soon extended to various branches of cell biology. Intracellular signaling pathways require adequate membrane organization, and the concept of membrane compartments provided the substrate to explain how dynamic processes could occur separately and simultaneously in spatio-temporally separateddiscrete domains.
In the central nervous system (CNS), most of the excitatory synapses occur on a specialized subcellular formation in neuronal arborizations, so-called dendritic spines. These spines constitute highly differentiated subcellular compartments of the receiving (postsynaptic) neuron, concentrating in a very small volume an abundant collection of neurotransmitter receptors, enzymes, scaffolding proteins, and cytoskeletal elements. Various neurotransmitter receptor systems harbored in the spine postsynaptic membrane have been shown to be associated with liquid-ordered, raft-type lipid domains [7,8,9,10]. These lipid domains not only harbor the receptor proteins but also recruit them from extra-synaptic regions; once receptors reach the active zone of the spine, the more rigid liquid-ordered domains contribute to reducing their lateral mobility [11,12]. These neurotransmitter receptor-containing lipid assemblies have dimensions in the order of nanometers, hence the designated term “nanodomains”. The surface of the mature spine, a morphological entity of ≤1 μm diameter, is thus sub-compartmentalized into much smaller parcels that contribute to the lateral heterogeneity of the postsynaptic membrane. This description of the dendritic spine is, however, fragmentary inasmuch it does not contemplate its highly dynamic nature, especially of some of its patchy nanodomains. Thus both stable and dynamic platforms coexist in the spine postsynaptic complex, and the dynamic range of the two combined is quite ample, from a few milliseconds to days [13], providing spatio-temporal stability to some of its components while allowing fast turnover and agile redistribution in others.
Pre- and postsynaptic membranes share the property of organizing their constituents in the form of nanodomains characterized by high molecular densities and transient duration. A characteristic feature of this organizational principle is that it is asymmetric (see review in [14]), in accordance with the vectorial nature of the synapse.
In this review, we focus on the structural and functional roles played by lipids and proteins in such nanodomains, discussing the significance of compartmentalization for efficient synaptic signaling. Examples are provided on how the glutamatergic and the cholinergic neurotransmitter receptors, two well characterized fast operating ligand-gated ion channels, and the endocannabinoid system, a typical metabotropic second-messenger G-protein coupled receptor (GPCR), tend to localize in clustered structures, the cholinergic and the cannabinoid receptors often overlapping anatomically, and share common roles in several physiological and pathological processes that exploit and rely on membrane nanodomains.

2. Dendritic Spines, Discrete Subcellular Compartment of the Neuronal Membrane

Dendritic spines constitute fundamental units of information endowed with processing synaptic (mostly excitatory) chemical neurotransmission in the mammalian brain. Spines compartmentalize biochemical and electrical signals, thus modulating the functional properties of synapses. Based on the relative length of their neck and the diameter of their head, dendritic spines have been arbitrarily classified into five categories: mushroom, thin, stubby, filopodia, and bifurcation- or cup-shaped spines [15]. However, this represents a fragmentary, anatomically static view of the dendritic spine: in fact, they also have a dynamic dimension, allowing them to modify their size and shape within seconds to minutes and undergo more lasting changes in the time scale of hours to days. In mature neurons spine motility declines [13]. Spine head volumes range from 0.01 to 1 μm3, and spine necks vary between 50 and 500 nm in diameter and up to 3 μm in length [16,17,18]. The size of the spine head has been shown to correlate with the size of the postsynaptic density (PSD) a characteristic morphological differentiation of the postsynapse [16,17,18], and with the amplitude of the excitatory postsynaptic current (EPSC) [19,20].
One of the most important functions of the brain is enabling the neural activity generated by an experience to modify neural circuit functionality in a process referred to as synaptic plasticity. This property of individual synapses affects higher brain functions like thought, feeling, and behavior [21]. Morphological correlates of synaptic plasticity at the dendritic spine level have been described, involving the remodeling of the PSD to regulate the number of intervening receptors and scaffolding proteins. The so-called long-term potentiation (LTP) and long-term depression (LTD), two mechanisms that have been thoroughly studied in the mammalian hippocampus, are induced by recent patterns of activity that strengthen or depress synapse activity, respectively [22] (Figure 1).
These functional readouts have straightforward morphological correlates: spine head enlargement in LTP [23,24] and conversely, shrinkage of the spine head and increased spine loss, as observed in hippocampal synapses after electrical induction of LTD [25,26]. The width of the spine neck has been associated with the degree of compartmentalization of synaptic signals [27], and its length with the modulation of synaptic membrane tension [28]. Upon induction of LTP, spine necks become shorter and wider [27] (Figure 2).
Compartmentalization of dendritic spines spatially constrains the diffusion of second messenger molecules to small volumes [29], thus circumscribing biochemical cascades and regulating information processing in individual synapses to smaller regions thereof. For example, presynaptic stimulation can elicit Ca+2 transient currents that are confined to hot spots of activity in spines [30,31]. The biochemical compartmentalization of the membrane requires membrane-bound enzymes, lipids and proteins to be organized into delimited regions of varying size and composition [32]. In addition to the delimitation in space, membrane domains vary in duration, thereby intervening in the timing of biochemical events at the synapse. This spatio-temporal organization extends to the coordination between pre- and postsynaptic components of the synaptic micro-cosmos (Figure 3).

3. Cannabinoids and Cannabinoid Receptors

The endocannabinoid system (ECS) is involved in almost all aspects of mammalian physiology and pathology. The ECS consists of genes encoding cannabinoid receptors (CBRs), endogenous cannabinoid ligands (endocannabinoids, eCBs), and the enzymes involved in their synthesis and degradation. Cannabinoids can be grouped into three major subclasses: phytocannabinoids, which are extracted from plants; eCBs, produced by live organisms, and synthetic cannabinoids that share some or all of the physiological properties of either phytocannabinoids and/or eCBs.
eCBs are lipidic, arachidonic acid-containing neurotransmitters (or messengers) that are synthesized de novo by phospholipase action after hydrolyzing the lipid precursors from the cellular membrane [33]. They are produced on demand when needed and are not usually stored in vesicles like classical neurotransmitters [34]. The synthesis of eCBs takes place upon postsynaptic depolarization through Ca2+ influx and activation of Gq-protein. The heterotrimeric G protein Gq stimulates eCB production through phospholipase C and D activation. After being synthesized in postsynaptic neurons, eCBs are released to act on CBRs expressed in presynaptic and/or nearby neurons.
Arachidonoyl ethanolamine (anandamine, AN), the first eCB discovered [35], is an endogenous lipid neurotransmitter derived from the polyunsaturated fatty acid arachidonic acid. AN is more soluble in aqueous media than arachidonic acid. AN acts as a partial agonist of CBR1 and CBR2 and as a full agonist of the vanilloid receptor 1 (VR1). The other eCB that acts as a full agonist of CBR1 and CBR2 is 2-arachidonyl glycerol (2-AG). Additionally, eCBs can activate other “non-CB” receptors, such as G protein-coupled receptor 55 (GPR55), peroxisome proliferator-activated receptors (PPARs) [36]. Although CBRs are mostly found on plasma membranes, their localization at intracellular compartments such as the endoplasmic reticulum (ER), endosomes, lysosomes, mitochondria and nuclei has also been demonstrated [37].
The two cannabinoid receptors identified to date have been termed CBR1 and CBR2. They share 44% amino acid sequence homology [38,39]. CBR1 constitutes the most abundant G-protein coupled receptor (GPCR) in the CNS [40], whereas CBR2 is localized preferentially in immune cells and peripheral tissues [39] and to a lesser extent in the brainstem, cortex, hippocampus and cerebellar neurons and microglia [41,42]. Some authors have suggested that VR1, also identified as the transient receptor potential cation channel subfamily V member 1 (TRPV1), should be classified as a cannabinoid 3-type receptor.
In the hippocampus CBR1s are mostly found in synapses that use γ-aminobutyric acid (GABA) as a neurotransmitter and mainly localized presynaptically, whereas CBR2s have a postsynaptic localization although their presence in presynaptic terminals cannot be excluded in some brain regions [43].
One of the physiological roles of CBR1s in the CNS is the regulation of synaptic transmission. eCBs attenuate presynaptic depolarization and subsequent neurotransmitter release [44]., thus mediating a retrograde regulation of this process. Upon activation, CBR1s repolarize the plasma membrane through the modulation of voltage-dependent ion channels that promote inward-rectifying K+ (Kir) and A-type K+ channels [45,46]. This is accomplished via the reduction in cAMP levels and PKA activation [47,48]. In addition, CBR1 activation can inhibit N-, P/Q- and L-type voltage-gated Ca2+ channels [49], thereby reducing Ca2+ influx and favoring repolarization of the presynaptic plasma membrane (see Figure 3).
Both CBR1 and CBR2 activation have been shown to favor phosphorylation and activation of the mitogen-activated protein (MAP) kinase cascade [50,51,52]. This metabolic cascade has an important role in the regulation of neuronal gene expression and synaptic plasticity. eCBs signaling is terminated by reuptake into both neurons and glial cells. Intracellular hydrolysis of AN and 2-AG is then catalyzed by the fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL) enzymes, respectively [53]. These lipid-derived molecules control the integration of signals generated by different neurotransmitters and synaptic neuromodulators that ultimately lead to diverse forms of homo-synaptic and hetero-synaptic meta-plasticity [54,55]. Thus, eCBs influence the flow of information that is key for the performance of many important high-order brain functions e.g., behavior [55,56,57].
Cannabinoids have been shown to alter the organization of the actin cytoskeleton in various cell types [58]. Njoo and collaborators highlighted the functional significance of CBR1 interactions with the Wiskott-Aldrich syndrome protein-family verprolin-homologous protein 1 (WAVE1)/SCAR1 complex [59]. These authors stated that the complex plays a key role in dynamically regulating the actin cytoskeleton in developing and adult neurons, thus contributing to the most salient functions of eCBs in the brain and spinal cord. Furthermore, they showed that cannabinoids structurally remodel dendritic spines by regulating the activity levels of WAVE1, and reported a novel function for WAVE1 in mediating inflammatory pain via structural and functional plasticity of spinal neurons [59].
Although CBR1s are mainly expressed presynaptically, several studies have indicated that CBR1 can be additionally localized in neuronal dendrites [60,61] colocalizing with the PSD protein PSD-95 in spines [59]. Rac1 activity decreases within minutes of CBR1 activation, limiting the conversion of G-actin to F-actin in dendritic spines of mature cortical neurons. This in turn produces a depletion of mature spines with the characteristic adult-type, elaborate mushroom morphology, which are believed to mediate increased synaptic efficacy such as observed in LTP [62]. In agreement with these findings, chronic treatment with the synthetic cannabinoid agonist WIN 55212-2 (WIN) reduced spine density in the nucleus accumbens (Nac) 24 h after the last injection [63]. Spiga and collaborators [64,65] also reported a decrease in spine density in the Nac (core) after a 1 h withdrawal period.
Activation of CBR1 in hippocampal neurons elicits a decrease in presynaptic F-actin and other cytoskeletal proteins, including ARPC2 and WASF1/WAVE1 that correlate with morphological changes (e.g., reduction in bouton size) [66], as seen at the postsynapse.

4. Cannabinoid Regulation of Synaptic Function and Compartmentalization

Neuronal circuits in hippocampus are fundamental in memory processes. Hippocampal pyramidal neurons perform important functions related to brain rhythms. Pyramidal cells are controlled by GABAergic interneurons, key for spatial memory formation. Potassium channels of the Kv7 type (KCNQ-Kv7) are characteristically located at the axon initial segment, where they modulate the membrane potential of the neuron and its cell firing. Oriens lacunosum moleculare interneurons, which are selectively active during theta oscillations, generate feedback inhibition of pyramidal neurons. In these somatostatin-positive interneurons, KCNQ-Kv7s are found predominantly in dendrites. Kv7 channel activity is upregulated following induction of presynaptic LTD. This depression of neuronal excitability involves eCB biosynthesis and depends on CBR1s, whereas eCBs acting directly on Kv7.2/3 channels, without participation of CBR1s, depress intrinsic neuronal excitability [67].

5. Cholinergic Signaling Contribution to Glutamatergic Receptor Compartmentalization

The paradigm excitatory neurotransmission in brain is mediated by glutamatergic synapses. These, in turn, are modulated by activation of presynaptic and postsynaptic neuronal nAChRs [68,69]. In the human brain, heteromeric α4β2 and homomeric α7 nAChRs constitute the most abundant subtypes of nAChRs, whereas other combinations are present in smaller numbers, and usually exhibit a more limited anatomical distribution, restricted to specific brain regions [70,71]. nAChRs are located on neurons where they regulate neurotransmitter release, cell excitability, and neuronal integration [72,73,74]. Activation of nAChRs can promote the release of glutamate (Glu), the main excitatory transmitter in mammalian brain, in a Ca2+-dependent manner. Presynaptic α4β2 or α7 nAChRs activation depolarizes hippocampal interneurons and indirectly affects the release of neurotransmitters like Glu and γ-aminobutyric acid (GABA) by activating voltage-gated Ca2+ channels [75]. Additionally, α7 nAChRs localized on brain glutamatergic terminals increases glutamate release upon activation [76] (Figure 1).
There is also evidence that the α4β2 and α7 subtypes of neuronal nAChRs contribute to changes in synaptic architecture, promoting glutamatergic synaptogenesis [77,78]. Postsynaptically located α7 nAChRs cause fast inward cationic currents in the neocortex and hippocampus. At this location α7 nAChRs have also been proposed to modulate synaptic potentiation independently of fast excitatory transmission [79] by recruiting GluA1 receptors from the surface pool of mobile extrasynaptic receptors and contributing to their stabilization in the spine via formation of receptor nanoclusters.. This stabilization requires functional PSD-95 scaffold protein family members that compartmentalize GluA1 macromolecules and help reduce their local mobility in conjunction with the actin submembrane meshwork. As a result, synapses enhance their signaling capacity, as reflected in the appearance of larger miniature excitatory postsynaptic currents (mEPSCs). This mechanism of GluA1 activation via α7 nAChRs is independent of fast excitatory transmission mediated by α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) or N-Methyl-D-aspartic acid (NMDA) receptors (NMDARs, see Figure 1).
Nicotine can bind to nAChRs on dopaminergic neurons [80] or on nerve terminals of GABAergic and glutamatergic neurons that project on the dopaminergic neurons in the ventral tegmental area (VTA). Neurons from the VTA project to neurons in the nucleus accumbens (NAc) and are involved in the processing of reward signals that motivate behavior. Thus, activation of nAChRs induces the release of dopamine in the mesolimbic dopaminergic pathway, involving the cholinergic pathways in different brain reward mechanisms [81,82]. Dopaminergic signaling in the hippocampus, the amygdala and the prefrontal cortex help to create emotional associations with rewards. Hence, activation of nAChRs in these areas contributes to the regulation of dopaminergic homeostasis and the rewarding and psychostimulant effects of addictive drugs, including nicotine [83,84]. α4β2 nAChRs have been linked to nicotine reward since studies performed in rats showed that their blockage with antagonists significantly inhibited nicotine self-administration [85,86] while genetic deletion of α4 or β2 subunits prevented nicotine-induced increase in NAc dopamine levels [86,87,88]. Other nAChR subunits, including α3, α6, α7 and β3, have been reported to mediate dopamine release after nicotine induction [89].
Crosstalk between the dopaminergic and glutamatergic systems enables them to initiate and organize normal behavior [90]. NMDARs can act as a scaffold to recruit laterally diffusing dopamine D1 receptors (D1Rs) to spines. The activation of these NMDARs alters the topography and movement of D1Rs by trapping them in dendritic spines [91]. D1Rs selectively interact with the NR1 subunit of the NMDAR through its C-terminal tail to form dimeric hetero-complexes. Induction of LTP in the striatum requires activation of D1Rs since antagonizing these receptors blocks NMDAR–dependent LTP; while in the cortex, working memory is altered by this antagonism [92,93,94]. Activation of D1Rs via DA release, caused e.g., by cholinergic signaling activation, can then recruit D1R-NMDAR complexes in a regulated manner [95].
Cholinergic signaling pathways thus contribute to sub-compartmentalize glutamatergic neurotransmission, spatially restricting its sphere of action. More generally, the ubiquitous distribution of nicotinic receptors in brain enables them to regulate many important high-level cognitive functions such as attention, working memory, learning processes [96], cognitive flexibility [97] and social interactions [98]. They are also involved in addiction and dependence [81,82].

6. Cannabinoids and Nicotinic Receptors

In addition to the well characterized activation of CBRs, cannabinoids also exert receptor-independent and lipid bilayer-mediated actions on many membrane proteins, enzymes, and transporters [99,100]. There is evidence of overlapping distribution of CBRs and nAChRs in many brain structures [101]. At the level of brain reward pathways, for instance, the endocannabinoid and nicotinic cholinergic systems interact bidirectionally, having an important role in the modulation of drug dependence [101]. In animal studies, the primary psychoactive ingredients in cannabis and tobacco, delta-9-tetrahydrocannabinol (Δ9-THC) and nicotine, respectively, show overlapping pharmacological effects, including induction of antinociception, hypothermia, rewarding effects, dependence, and impairment of locomotion [102,103,104,105,106,107]. Additionally, Δ9-THC and nicotine potentiate their anxiolytic effects in a synergistic manner when administered together at sub-threshold doses [108,109,110,111]. Chronic administration of nicotine in rats increases AN levels in the limbic forebrain and of both AN and 2-AG in the brainstem [89]. Reciprocally, Δ9-THC and synthetic cannabinoids modulate cholinergic neurotransmission in the hippocampus [112,113,114], supporting the notion that nicotinic and endocannabinoid systems interact in the brain reward pathway. This information opens a window for therapeutic intervention in nicotine and cannabinoid dependence [115,116].
Other experimental demonstrations of how cannabinoids and nicotinic receptors interact can be found in the direct actions of endocannabinoids on α7 nAChR in Xenopus oocytes [117,118]. In this experimental model system, AN reversibly inhibited nicotine-induced currents in a concentration-dependent manner without significantly affecting its half maximal effective concentration (EC50) value, indicating that AN acts as a noncompetitive antagonist on α7 nAChRs. In agreement with these observations, Frey and coworkers reported that AN does not inhibit specific [3H]nicotine binding in human frontal cortex membranes [119], providing supporting evidence for the noncompetitive action of AN on nicotinic receptors. A direct inhibitory action of AN on nAChRs is suggested by the lack of effect on CBR1, CBR2, or enzymes involved in AN metabolism. It must be noted however that the synthetic cannabinoids WIN55,212-2 and THC do not alter α7 nAChR function. Instead, AN markedly inhibits the peak amplitudes of α4β2 nAChR-mediated currents to about 30% of control levels in SH-EP1 cells [120]. In addition, studies performed on myenteric neurons have shown that AN can reduce the amplitudes of nicotine-induced inward currents [99,121].
The lipophilic nature of eCBs facilitates their incorporation into biological membranes, and their accumulation has been shown to modulate the action of alcohol and volatile anesthetics on α7 nAChRs [122,123]. These pharmacological effects suggest that CBs may alter the bulk physicochemical properties of the plasma membrane. Thus, the two types of endogenous neurotransmitters, eCB and acetylcholine can directly or indirectly (following binding to hydrophobic sites on the receptor surfaces) modulate the function of these receptors in brain and exert modulatory crossover effects on each other.
In the activated spine, Ca2+ entry through postsynaptic α7 nAChRs [70] or NMDA receptors and/or voltage-gated Ca2+ channels [124] can promote the activation of multiple signaling pathways with specific spatio-temporal patterns that orchestrate and regulate different aspects of cytoskeletal dynamics in the stimulated spines. Within the spine, Ca2+ binds to calmodulin (CaM), a Ca2+-binding protein which subsequently activates Ca2+/CaM-dependent kinases and phosphatases such as CaMKII and calcineurin [125,126]. CaMKII activates small GTPases and these in turn further modulate several downstream kinases [127] that have the ability to activate many ABPs, including Cofilin and Arp2/3, two proteins that play essential roles in actin remodeling [128,129]. Thus cannabinoids, through rac1/WAVE 1 modulation and α7 nAChR activation, contribute to actin structural remodeling of dendritic spines. Whereas activation of Rac1/WAVE1 induces a depletion of mushroom type spines, α7 nAChR activation contributes to the formation and maturation of dendritic spines (Figure 1).

7. Importance of Lipids in Dendritic Spine Compartmentalization

Lipids play a central role in the formation and shaping of membranes [130,131]. More than 40 different lipids modulate signaling and dynamically influence membrane geometry of synapses [132,133,134,135]. As in other biological membranes, the major classes of lipids found at synaptic plasma membranes are cholesterol, sphingolipids and glycerol backbone-based phospholipids [136]. Of these, cholesterol and saturated sphingolipids and phospholipids are particularly enriched in membrane domains [137], and the cholesterol concentration in brain accounts for ~25% of the human body’s total content.
The neutral lipid cholesterol plays a key role in receptor clustering and compartmentalization [77,79]. The amount of cholesterol and its topographical distribution at the plasma membrane have a profound influence on the physicochemical properties of the plasmalemma, and these, in turn, modulate the functional activity of the synapse. The most general effect of cholesterol is exerted on the physical properties of the bulk membrane. Membrane cholesterol concentrations above a certain threshold level promote the formation of liquid-ordered (Lo) domains [138] that are thicker than the rest of the membrane and exhibit unique biophysical properties. These ordered lipid nanodomains (“lipid rafts”) mediate trafficking, signal transduction, and membrane protein activity in essentially all eukaryotic plasma membranes [139].
Cholesterol increases membrane rigidity, thickness and tension that influences cell signaling and domain formation [140,141]. The concept of lipid domains entails not only the existence of lateral heterogeneities in the plane of the membrane, but also the dynamic exchange of molecular components within and between compartments. For instance cholesterol flip-flops between the outer and inner leaflets of the membrane bilayer, and also diffuses within a given liquid-ordered lipid domain and between different lipid domains in inter-domain exchange processes which are quite heterogeneous, differing considerably in their kinetic parameters [142].
At the synapse, cholesterol interacts with several neurotransmitter receptors through consensus linear binding sequences like the so-called cholesterol recognition/interaction amino acid consensus motifs (CRAC and its mirror image CARC [143]. These consensus domains have been proposed to facilitate membrane protein incorporation into cholesterol-rich domains in a great variety of membrane proteins, including the superfamily of pentameric ligand-gated ion channels (pLGICs) and the superfamily of G-protein coupled receptors (GPCRs). The prototypic LGIC, the nAChR, exhibits a CRAC motif adjacent to the transmembrane helix M1, and a CARC sequence on the M4-facing surface of M1 adjacent to one of the proposed cholesterol-binding cavities [143,144]. Likewise, the transmembrane helix 7 of human cannabinoid receptor 1 (CBR1) displays a CRAC sequence [145]. A cholesterol molecule was recently identified in a crystal structure [146] and a cryo-electron microscopy (EM) structure [147] of the CBR1.
Sphingolipids participate as plasma membrane lipids and signaling molecules such as ceramide, sphingosine, and sphingosine-1-phosphate that are produced after the metabolism of plasma membrane sphingolipids [148]. Among sphingolipids, sphingomyelins are enriched in brain membranes. The elimination of dendritic spines upon reduction of cholesterol and sphingomyelin levels was described almost 20 years ago [149], highlighting the importance of these lipids for neuronal communication. Ceramide promotes spine maturation by contributing to the transformation of dendritic filopodia to mature spines [150].
Phosphoinositides (PIPs) are important players in postsynaptic excitability since they have an exceptional high rate of metabolic turnover and compartmentalization [151]. There are multiple enzymes at dendritic spines that interconvert different PIPs contributing to the dynamic role of lipids in the plasma membrane. Phosphatidylinositol (4,5) diphosphate (PIP2) is converted to phosphatidylinositol (3,4,5) triphosphate (PIP3) by phosphatidylinositol-4,5-bisphosphate 3-kinase. PIP3 content at the spines is higher than that in dendritic shafts under basal conditions. Furthermore, upon glutamate stimulation PIP3 redistributes contributing to the formation of fine spinules projecting from spines [152,153]. Additionally, PIP3, because of its capacity regulate the activity of multiple Rho GTPase effectors [154], has also been implicated in the interaction of membrane-cytoskeleton crosstalk at spines, and is able to regulate the Akt-mTOR pathway to participate in dendritic spine morphogenesis [155,156]. The PIP2-clustering molecule myristoylated alanine-rich C kinase substrate (MARCKS) reversibly sequesters PIP2 on the plasma membrane, upon local increases in intracellular calcium [157]. MARCKs contributes to spine morphogenesis by promoting the transition from immature dendritic spines to larger and more stable mushroom-shaped spines by controlling actin cytoskeleton [158]. Moreover, association of MARCKS to cholesterol at the membrane is necessary for its ability to crosslink F-actin [158]. In addition, PIP3 contributes to the accumulation of PSD-95 at spines whereas conversion of PIP2 by phospholipase C favors synaptic actin depolymerization and PSD-95 degradation, thus contributing to spine remodeling [159].
Among other factors, actin dynamics is modulated by membrane lipids. A reduction in membrane cholesterol levels produces a rapid collapse of spine morphological integrity associated with redistribution of F-actin from the spine proper to the dendritic shaft [149]. Sphingolipids also play a relevant role in the spine plasma membrane-actin cytoskeleton crosstalk. Sphingomyelins modulate membrane binding and activity of the Rho GTPases, key regulators of the actin cytoskeleton in the synapse. Accumulation of sphingomyelins at postsynaptic membranes, as observed in a Niemann-Pick disease type A mouse model defective in acid sphingomyelinase, induces a reduction of metabotropic glutamate receptors that impairs the membrane attachment of RhoA and its effectors ROCK (RhoA kinase) and profilin IIa. This impairment results in the diminution of F-actin content, ultimately reducing spine number and size [160]. The conversion of sphingomyelin to ceramide at the plasma membrane is catalyzed by neutral sphingomyelinase-2 [161]. This enzyme can, in turn, modulate spine actin cytoskeleton. Activation of the neutral sphingomyelinase restores F-actin content of dendritic spines by enhancing the RhoA pathway in mice defective of the acid sphingomyelinase, which present high sphingomyelin synaptic levels [160]. Conversely, inhibition of the sphingomyelinase decreases the abnormally high levels of F-actin in spines of neurons in mice lacking Wiskott–Aldrich syndrome protein interacting protein (WIP) [162]. Thus, through various protein and lipid interactions, actin promotes compartmentalization of the plasma membrane and has an important role in the modulation of neuronal strength.

Impact of the Lipid Microenvironment on nAChRs and CBRs

The best documented example of the influence that the lipid microenvironment and cholesterol exert on the topography and function of a neurotransmitter receptor is provided by the paradigm rapid LGIC, the nAChR (see reviews in [163,164]). Changes in cholesterol levels alter the translational mobility of the receptor in the plane of the plasma membrane, as measured by fluorescence recovery after photobleaching and fluorescence correlation spectroscopy [165] and single-molecule localization microscopy [166,167]. Cell-surface trafficking of nAChRs is dependent on cholesterol metabolism [168,169]. Pharmacological long-term inhibition of cholesterol biosynthesis by the statin lovastatin differentially augments cell-surface levels of α4β2 and α7 nAChRs in neurites and soma of rat hippocampal neurons [169]. Misbalances in brain cholesterol homeostasis affect cholinergic signaling involving α4β2 and α7 nAChRs and indirectly impact on the number and distribution of other neuroreceptors at dendritic spines, with important consequences for brain function. Sphingolipids are also necessary for nAChR export in the early secretory pathways [170]. Thus, any modification in sphingolipid levels will impact on nAChR expression.
CBRs are also influenced by the lipid composition of the cell membrane [171]. In particular, cholesterol has been reported to negatively modulate the activity of CBR1 in nerve cells [172] and a specific cholesterol binding site has been described in the CBR1 molecule [145]. Likewise, cholesterol has recently been reported to increase the basal activation levels of the CBR2 receptor by exerting an allosteric effect on the intracellular regions of the receptor [171]. Considering the retrograde regulatory effects of CBRs on neurotransmitter release [44] lipid compartmentalization by membrane cholesterol is of considerable importance as it provides additional regulatory check points of neurotransmitter receptor function.
Furthermore, the lipophilic endogenous neurotransmitter AN exhibits selectivity for cholesterol and ceramide amongst membrane lipids. Through the establishment of a hydrogen bond, AN forms a molecular complex with cholesterol inside the plasma membrane, and has been proposed to interact with CBR1 via the cholesterol-recognition motifs CRAC or CARC [173]. AN binding to ceramide blocks the degradation pathway of both lipids [173].

8. Contribution of Actin Dynamics to the Compartmentalization of the Dendritic Spine

The eukaryotic plasma membrane is connected to the intracellular milieu via many adapter proteins [174]. The plasma membrane domains are associated with intracellular structures, and the submembrane cytoskeletal meshwork is one of the important elements that regulate cell mechanics and morphology, contributing to delimiting the extension of membrane domains. The actin cytoskeleton modulates plasma membrane dynamics and introduces barriers to the diffusion of membrane proteins, with the formation of micrometer-sized corrals and smaller domains where membrane proteins can experience anomalous diffusion [175,176,177,178]. These actin meshwork-driven compartments play an important role in the dendritic spine. For instance, disruption of PDZ-containing scaffolds (PDZ domains are protein–protein interaction modules that recognize specific C-terminal sequences) or of actin filaments in chick ciliary ganglion neurons has been shown to increase the mobility of α7 nAChRs [179]. In contrast, making the sub-membrane cortical actin network more stable by cholesterol depletion [180] reduces receptor mobility. Cholesterol depletion diminishes nAChR mobility at the cell surface; mobility can be partially restored upon treatment with Latrunculin A [165].
Actin polymerization and depolymerization dynamics are active modifiers of presynaptic morphology, with important functional implications [181,182]. Thus, the actin cytoskeleton provides the adequate scaffold for synaptic vesicles and the active zone [183,184]. Changes in the size of presynaptic terminals have been correlated with the postsynaptic response [185,186,187] and thus with synapse strength [187,188,189,190]. At the postsynapse, actin filament dynamics provide a framework for the formation of transient protrusions and modifications of dendritic spine morphology by interacting with a variety of actin-binding proteins (ABPs) [191]. These ABPs are important in actin dynamics, including actin polymerization-depolymerization, nucleation, branching, capping, cross-linking and trafficking [192].

9. Nanodomain Cluster Organization Emerges as a Common Organizing Principle at CNS Synapses

Neuronal activity regulates receptor fluxes at the PSD through the interaction of receptors with scaffolding molecules and lipids [193]. The PSD scaffolding proteins are distributed heterogeneously and form nanodomains within the synapses, termed sub-synaptic domains (SSD) [194]. The characteristic output of LTP induction, spine enlargement, is paralleled by the increase in the number of PSD-95 (an SSD-resident protein) copies [195]. This augmentation in PSD-95 biosynthesis occurs within hours of LTP induction [196].
During development, proteins organized in nanoclusters at the pre- and postsynapse control the differentiation of the dendritic spine. The cholesterol-binding protein TSPAN5, which belongs to the tetraspanin superfamily, is localized postsynaptically in pyramidal excitatory neurons, and is the master controller of spine maturation. Control is exerted by promoting the clustering of the postsynaptic molecule neuroligin-1 at the spine surface [197], which in turn recognizes and binds to presynaptic neurexins. TSPAN5 knockout mice have an extremely low number of spines, reinforcing the view that this tetraspanin plays a key role in excitatory synapse maturation.
Neurotransmitter receptors are also organized into SSDs at the postsynaptic plasma membrane [198]. Neurotransmitter receptor domains localized in SSD domains exhibit diffusion-limited lateral mobility relative to receptors localized in non-synaptic areas, and the two pools are in active exchange [199]. Receptors in either pool are removed from the plasma membrane via metabolic turnover [200]. The maintenance of a homeostatic balance between these two mechanisms determines the number of active receptors at the postsynaptic membrane.
At excitatory glutamatergic synapses, the lateral exchange of receptors between extrasynaptic and synaptic areas is responsible for neuronal plasticity [193,201,202]. In the CNS, endocytosis occurs predominantly, if not exclusively, in extrasynaptic areas [203,204,205], and hence receptors have to diffuse out of the PSD to be internalized [206]. In turn, mobilization of receptors from the spine shaft to the PSD partly depends on the morphology of the spine. As the spine neck diameter diminishes, it is energetically and mechanically more costly to transport endosomal vesicles from the ER outposts in the dendrite to the PSD through the spine actin meshwork [207]. Dendritic ER and Golgi outposts are well documented in rodent hippocampal neurons, particularly at dendritic branching points [208] and high-pressure freezing fixation (without chemical fixation) combined with transmission electron microscopy has revealed smooth ER and Golgi outposts inside dendritic spines [209] (Figure 4).
Formation of cell-surface receptor nanoclusters is favored when stabilizing interactions preclude lateral diffusion of individual receptor molecules [210,211]. Upon disruption of the cytoskeleton, α7 nAChR mobility and cell surface expression are hampered, but not the receptors’ ability to form clusters, suggesting tethering mechanisms [212].
Perisomatic GABAergic terminals from CBR1-positive basket cell interneurons produce stronger endocannabinoid-mediated synaptic plasticity than that mediated by dendritic synapses [213]. Consistent with these observations, studies combining stochastic optical reconstruction microscopy (STORM) superresolution optical microscopy with whole-cell patch-clamp electrophysiological recordings found that GABAergic interneurons projecting axon terminals on perisomatic synapses possess larger synaptic boutons, harboring higher numbers of CBR1s, higher receptor/effector ratio, and more complex active zones than dendritic spine-projecting interneurons, whereas nanoclusters of the presynaptic active zone-resident protein Bassoon were more abundant in dendritic synapses [214]. The CBR1/Bassoon ratio was 46% higher within 100 nm of the Bassoon nanocluster edges in perisomatic boutons than in dendritic nerve endings. These authors also found that chronic Δ9-THC administration, which diminishes cannabinoid efficacy on GABA release, markedly downregulated CBR1s. Complete receptor recovery occurred only weeks after the cessation of Δ9-THC treatment. Dudok and coworkers hypothesized that the nanoscale receptor/effector ratio and coupling distance, rather than the number of receptors, determines cannabinoid signaling and neurotransmitter release [214].
Recently, Goncalves and coworkers showed that NMDARs form a single SSD located mainly at the center of the PSD, whereas AMPARs were segregated into several SSDs surrounding the NMDAR central nanocluster and mGluR5 presents a more diffuse distribution [215]. This observation is in agreement with the notion that NMDARs are less mobile than AMPARs at mature synapses [202,216] (Figure 4B). As for CBRs, CBR1 colocalizes with PSD-95 in spines [59]. The predominantly presynaptic CBR1s have been reported to be uniformly distributed at the nanoscale level [214]. However, as is the case with the nAChR, changes in cholesterol/sphingolipid ratio and actin stability can alter the distribution of CBR1s. In addition, a reduction in the number of cell-surface CBR1s via endocytic internalization [47,217] or through the action of β-arrestins [218,219] can be induced after long agonist exposure. Binding of β-arrestins uncouples G-proteins and stimulates receptor internalization and β-arrestin mediated signaling [218,220].

10. Concluding Remarks

Compartmentalization provides the plasma membrane with discrete 2-D platforms where biochemical reactions or signaling mechanisms can be simultaneously executed with maximal efficacy, without interference from competing processes. The dendritic spine, a highly differentiated subcellar compartment distinct from the rest of the neuronal membrane, maximizes its operational capacity by incorporating nanodomain compartments at its surface. One of the mechanisms exploited by the dendritic spine to parcel out its cell-surface membrane is through the formation of reversible biological barriers via the actin meshwork. This sub-membrane network dynamically interacts with the plasmalemma, creating barriers that transiently hinder protein translational diffusion, thereby secluding molecules from the rest of the bilayer and transiently retaining them in nanodomain compartments.
Advances in the field of dendritic spine compartmentalization should provide important clues about the mechanisms involved in the regulation and function of neurotransmitter receptors and their interplay with scaffolding proteins, lipids, and enzymes at brain synapses in health and disease.

Author Contributions

F.J.B. and A.S.V. conceived the work, searched the literature, and wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

No funding is associated with this review.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

Figures were created using Servier Medical Art Commons Attribution 3.0 Unported License. (http://smart.servier.com). Servier Medical Art by Servier is licensed under a Creative Commons Attribution 3.0 Unported License.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kusumi, A.; Shirai, Y.M.; Koyama-Honda, I.; Suzuki, K.G.N.; Fujiwara, T.K. Hierarchical organization of the plasma membrane: Investigations by single-molecule tracking vs. fluorescence correlation spectroscopy. FEBS Lett. 2010, 584, 1814–1823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kusumi, A.; Fujiwara, T.K.; Tsunoyama, T.A.; Kasai, R.S.; Liu, A.-A.; Hirosawa, K.M.; Kinoshita, M.; Matsumori, N.; Komura, N.; Ando, H.; et al. Defining raft domains in the plasma membrane. Traffic 2020, 21, 106–137. [Google Scholar] [CrossRef]
  3. Jacobson, K.; Liu, P. Complexity Revealed: A Hierarchy of Clustered Membrane Proteins. Biophys. J. 2016, 111, 1–2. [Google Scholar] [CrossRef] [Green Version]
  4. Simons, K.; Ikonen, E. Functional rafts in cell membranes. Nature 1997, 387, 569–572. [Google Scholar] [CrossRef] [PubMed]
  5. Jacobson, K.; Dietrich, C. Looking at lipid rafts? Trends Cell Biol. 1999, 9, 87–91. [Google Scholar] [CrossRef]
  6. Simons, K.; Toomre, D. Lipid rafts and signal transduction. Nat. Rev. Mol. Cell Biol. 2000, 1, 31–39. [Google Scholar] [CrossRef]
  7. Brusés, J.L.; Chauvet, N.; Rutishauser, U. Membrane lipid rafts are necessary for the maintenance of the (alpha)7 nicotinic acetylcholine receptor in somatic spines of ciliary neurons. J. Neurosci. 2001, 21, 504–512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Hayashi, T.; Su, T.-P. Sigma-1 receptors at galactosylceramide-enriched lipid microdomains regulate oligodendrocyte differentiation. Proc. Natl. Acad. Sci. USA 2004, 101, 14949–14954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Allen, J.A.; Halverson-Tamboli, R.A.; Rasenick, M.M. Lipid raft microdomains and neurotransmitter signalling. Nat. Rev. Neurosci. 2007, 8, 128–140. [Google Scholar] [CrossRef] [PubMed]
  10. Egawa, J.; Pearn, M.L.; Lemkuil, B.P.; Patel, P.M.; Head, B.P. Membrane lipid rafts and neurobiology: Age-related changes in membrane lipids and loss of neuronal function. J. Physiol. 2016, 594, 4565–4579. [Google Scholar] [CrossRef] [Green Version]
  11. Nagappan, G.; Lu, B. Activity-dependent modulation of the BDNF receptor TrkB: Mechanisms and implications. Trends Neurosci. 2005, 28, 464–471. [Google Scholar] [CrossRef] [PubMed]
  12. Fernandes, C.C.; Berg, D.K.; Gómez-Varela, D. Lateral mobility of nicotinic acetylcholine receptors on neurons is determined by receptor composition, local domain, and cell type. J. Neurosci. 2010, 30, 8841–8851. [Google Scholar] [CrossRef] [Green Version]
  13. Dunaevsky, A.; Tashiro, A.; Majewska, A.; Mason, C.; Yuste, R. Developmental regulation of spine motility in the mammalian central nervous system. Proc. Natl. Acad. Sci. USA 1999, 96, 13438–13443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Heine, M.; Holcman, D. Asymmetry between Pre- and Postsynaptic Transient Nanodomains Shapes Neuronal Communication. Trends Neurosci. 2020, 43, 182–196. [Google Scholar] [CrossRef] [PubMed]
  15. Hering, H.; Sheng, M. Dentritic spines: Structure, dynamics and regulation. Nat. Rev. Neurosci. 2001, 2, 880–888. [Google Scholar] [CrossRef]
  16. Trommald, M.; Hulleberg, G. Dimensions and density of dendritic spines from rat dentate granule cells based on reconstructions from serial electron micrographs. J. Comp. Neurol. 1997, 377, 15–28. [Google Scholar] [CrossRef]
  17. Arellano, J.I.; Benavides-Piccione, R.; Defelipe, J.; Yuste, R. Ultrastructure of dendritic spines: Correlation between synaptic and spine morphologies. Front. Neurosci. 2007, 1, 131–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Harris, K.M.; Stevens, J.K. Dendritic spines of CA 1 pyramidal cells in the rat hippocampus: Serial electron microscopy with reference to their biophysical characteristics. J. Neurosci. 1989, 9, 2982–2997. [Google Scholar] [CrossRef] [PubMed]
  19. Noguchi, J.; Nagaoka, A.; Watanabe, S.; Ellis-Davies, G.C.R.; Kitamura, K.; Kano, M.; Matsuzaki, M.; Kasai, H. In vivo two-photon uncaging of glutamate revealing the structure-function relationships of dendritic spines in the neocortex of adult mice. J. Physiol. 2011, 589, 2447–2457. [Google Scholar] [CrossRef] [PubMed]
  20. Matsuzaki, M.; Ellis-Davies, G.C.R.; Nemoto, T.; Miyashita, Y.; Iino, M.; Kasai, H. Dendritic spine geometry is critical for AMPA receptor expression in hippocampal CA1 pyramidal neurons. Nat. Neurosci. 2001, 4, 1086–1092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Citri, A.; Malenka, R.C. Synaptic Plasticity: Multiple Forms, Functions, and Mechanisms. Neuropsychopharmacology 2008, 33, 18–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Bliss, T.V.P.; Cooke, S.F. Long-term potentiation and long-term depression: A clinical perspective. Clinics 2011, 66 (Suppl. 1), 3–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Matsuzaki, M.; Honkura, N.; Ellis-Davies, G.C.R.; Kasai, H. Structural basis of long-term potentiation in single dendritic spines. Nature 2004, 429, 761–766. [Google Scholar] [CrossRef] [PubMed]
  24. Lang, C.; Barco, A.; Zablow, L.; Kandel, E.R.; Siegelbaum, S.A.; Zakharenko, S.S. Transient expansion of synaptically connected dendritic spines upon induction of hippocampal long-term potentiation. Proc. Natl. Acad. Sci. USA 2004, 101, 16665–16670. [Google Scholar] [CrossRef] [Green Version]
  25. Zhou, Q.; Homma, K.J.; Poo, M. Shrinkage of dendritic spines associated with long-term depression of hippocampal synapses. Neuron 2004, 44, 749–757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Nägerl, U.V.; Eberhorn, N.; Cambridge, S.B.; Bonhoeffer, T. Bidirectional activity-dependent morphological plasticity in hippocampal neurons. Neuron 2004, 44, 759–767. [Google Scholar] [CrossRef] [Green Version]
  27. Tønnesen, J.; Katona, G.; Rózsa, B.; Nägerl, U.V. Spine neck plasticity regulates compartmentalization of synapses. Nat. Neurosci. 2014, 17, 678–685. [Google Scholar] [CrossRef] [Green Version]
  28. Vanderklish, P.W.; Edelman, G.M. Dendritic spines elongate after stimulation of group 1 metabotropic glutamate receptors in cultured hippocampal neurons. Proc. Natl. Acad. Sci. USA 2002, 99, 1639–1644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Tønnesen, J.; Nägerl, U.V. Dendritic Spines as Tunable Regulators of Synaptic Signals. Front. Psychiatry 2016, 7, 101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Müller, W.; Connor, J.A. Dendritic spines as individual neuronal compartments for synaptic Ca2+ responses. Nature 1991, 354, 73–76. [Google Scholar] [CrossRef] [PubMed]
  31. Guthrie, P.B.; Segal, M.; Kater, S.B. Independent regulation of calcium revealed by imaging dendritic spines. Nature 1991, 354, 76–80. [Google Scholar] [CrossRef] [PubMed]
  32. Honigmann, A.; Pralle, A. Compartmentalization of the Cell Membrane. J. Mol. Biol. 2016, 428, 4739–4748. [Google Scholar] [CrossRef] [PubMed]
  33. Correa, F.; Wolfson, M.L.; Valchi, P.; Aisemberg, J.; Franchi, A.M. Endocannabinoid system and pregnancy. Reproduction 2016, 152, R191–R200. [Google Scholar] [CrossRef] [PubMed]
  34. Di Marzo, V. The endocannabinoid system: Its general strategy of action, tools for its pharmacological manipulation and potential therapeutic exploitation. Pharmacol. Res. 2009, 60, 77–84. [Google Scholar] [CrossRef] [PubMed]
  35. Devane, W.A.; Hanus, L.; Breuer, A.; Pertwee, R.G.; Stevenson, L.A.; Griffin, G.; Gibson, D.; Mandelbaum, A.; Etinger, A.; Mechoulam, R. Isolation and structure of a brain constituent that binds to the cannabinoid receptor. Science 1992, 258, 1946–1949. [Google Scholar] [CrossRef]
  36. Rankin, L.; Fowler, C.J. The Basal Pharmacology of Palmitoylethanolamide. Int. J. Mol. Sci. 2020, 21, 7942. [Google Scholar] [CrossRef] [PubMed]
  37. Hebert-Chatelain, E.; Desprez, T.; Serrat, R.; Bellocchio, L.; Soria-Gomez, E.; Busquets-Garcia, A.; Pagano Zottola, A.C.; Delamarre, A.; Cannich, A.; Vincent, P.; et al. A cannabinoid link between mitochondria and memory. Nature 2016, 539, 555–559. [Google Scholar] [CrossRef]
  38. Matsuda, L.A.; Lolait, S.J.; Brownstein, M.J.; Young, A.C.; Bonner, T.I. Structure of a cannabinoid receptor and functional expression of the cloned cDNA. Nature 1990, 346, 561–564. [Google Scholar] [CrossRef]
  39. Munro, S.; Thomas, K.L.; Abu-Shaar, M. Molecular characterization of a peripheral receptor for cannabinoids. Nature 1993, 365, 61–65. [Google Scholar] [CrossRef]
  40. Herkenham, M.; Lynn, A.B.; Little, M.D.; Johnson, M.R.; Melvin, L.S.; de Costa, B.R.; Rice, K.C. Cannabinoid receptor localization in brain. Proc. Natl. Acad. Sci. USA 1990, 87, 1932–1936. [Google Scholar] [CrossRef] [Green Version]
  41. Núñez, E.; Benito, C.; Pazos, M.R.; Barbachano, A.; Fajardo, O.; González, S.; Tolón, R.M.; Romero, J. Cannabinoid CB2 receptors are expressed by perivascular microglial cells in the human brain: An immunohistochemical study. Synapse 2004, 53, 208–213. [Google Scholar] [CrossRef] [PubMed]
  42. Van Sickle, M.D.; Duncan, M.; Kingsley, P.J.; Mouihate, A.; Urbani, P.; Mackie, K.; Stella, N.; Makriyannis, A.; Piomelli, D.; Davison, J.S.; et al. Identification and functional characterization of brainstem cannabinoid CB2 receptors. Science 2005, 310, 329–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ong, W.Y.; Mackie, K. A light and electron microscopic study of the CB1 cannabinoid receptor in the primate spinal cord. J. Neurocytol. 1999, 28, 39–45. [Google Scholar] [CrossRef] [PubMed]
  44. Kano, M.; Ohno-Shosaku, T.; Hashimotodani, Y.; Uchigashima, M.; Watanabe, M. Endocannabinoid-mediated control of synaptic transmission. Physiol. Rev. 2009, 89, 309–380. [Google Scholar] [CrossRef]
  45. Deadwyler, S.A.; Hampson, R.E.; Mu, J.; Whyte, A.; Childers, S. Cannabinoids modulate voltage sensitive potassium A-current in hippocampal neurons via a cAMP-dependent process. J. Pharmacol. Exp. Ther. 1995, 273, 734–743. [Google Scholar] [PubMed]
  46. Vásquez, C.; Navarro-Polanco, R.A.; Huerta, M.; Trujillo, X.; Andrade, F.; Trujillo-Hernández, B.; Hernández, L. Effects of cannabinoids on endogenous K+ and Ca2+ currents in HEK293 cells. Can. J. Physiol. Pharmacol. 2003, 81, 436–442. [Google Scholar] [CrossRef]
  47. Howlett, A.C.; Breivogel, C.S.; Childers, S.R.; Deadwyler, S.A.; Hampson, R.E.; Porrino, L.J. Cannabinoid physiology and pharmacology: 30 years of progress. Neuropharmacology 2004, 47 (Suppl. 1), 345–358. [Google Scholar] [CrossRef]
  48. Hampson, R.E.; Evans, G.J.; Mu, J.; Zhuang, S.Y.; King, V.C.; Childers, S.R.; Deadwyler, S.A. Role of cyclic AMP dependent protein kinase in cannabinoid receptor modulation of potassium “A-current” in cultured rat hippocampal neurons. Life Sci. 1995, 56, 2081–2088. [Google Scholar] [CrossRef]
  49. Howlett, A.C.; Barth, F.; Bonner, T.I.; Cabral, G.; Casellas, P.; Devane, W.A.; Felder, C.C.; Herkenham, M.; Mackie, K.; Martin, B.R.; et al. International Union of Pharmacology. XXVII. Classification of cannabinoid receptors. Pharmacol. Rev. 2002, 54, 161–202. [Google Scholar] [CrossRef]
  50. Bouaboula, M.; Perrachon, S.; Milligan, L.; Canat, X.; Rinaldi-Carmona, M.; Portier, M.; Barth, F.; Calandra, B.; Pecceu, F.; Lupker, J.; et al. A selective inverse agonist for central cannabinoid receptor inhibits mitogen-activated protein kinase activation stimulated by insulin or insulin-like growth factor 1. Evidence for a new model of receptor/ligand interactions. J. Biol. Chem. 1997, 272, 22330–22339. [Google Scholar] [CrossRef] [Green Version]
  51. Bouaboula, M.; Poinot-Chazel, C.; Bourrié, B.; Canat, X.; Calandra, B.; Rinaldi-Carmona, M.; Le Fur, G.; Casellas, P. Activation of mitogen-activated protein kinases by stimulation of the central cannabinoid receptor CB1. Biochem. J. 1995, 312 Pt 2, 637–641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Derkinderen, P.; Ledent, C.; Parmentier, M.; Girault, J.A. Cannabinoids activate p38 mitogen-activated protein kinases through CB1 receptors in hippocampus. J. Neurochem. 2001, 77, 957–960. [Google Scholar] [CrossRef] [PubMed]
  53. Muccioli, G.G. Endocannabinoid biosynthesis and inactivation, from simple to complex. Drug Discov. Today 2010, 15, 474–483. [Google Scholar] [CrossRef] [PubMed]
  54. Alger, B.E. Endocannabinoid signaling in neural plasticity. Curr. Top. Behav. Neurosci. 2009, 1, 141–172. [Google Scholar] [CrossRef] [PubMed]
  55. Edwards, D.A.; Zhang, L.; Alger, B.E. Metaplastic control of the endocannabinoid system at inhibitory synapses in hippocampus. Proc. Natl. Acad. Sci. USA 2008, 105, 8142–8147. [Google Scholar] [CrossRef] [Green Version]
  56. Melis, M.; Greco, B.; Tonini, R. Interplay between synaptic endocannabinoid signaling and metaplasticity in neuronal circuit function and dysfunction. Eur. J. Neurosci. 2014, 39, 1189–1201. [Google Scholar] [CrossRef]
  57. Iremonger, K.J.; Wamsteeker Cusulin, J.I.; Bains, J.S. Changing the tune: Plasticity and adaptation of retrograde signals. Trends Neurosci. 2013, 36, 471–479. [Google Scholar] [CrossRef]
  58. Hohmann, T.; Feese, K.; Ghadban, C.; Dehghani, F.; Grabiec, U. On the influence of cannabinoids on cell morphology and motility of glioblastoma cells. PLoS ONE 2019, 14, e0212037. [Google Scholar] [CrossRef] [PubMed]
  59. Njoo, C.; Agarwal, N.; Lutz, B.; Kuner, R. The Cannabinoid Receptor CB1 Interacts with the WAVE1 Complex and Plays a Role in Actin Dynamics and Structural Plasticity in Neurons. PLOS Biol. 2015, 13, e1002286. [Google Scholar] [CrossRef] [Green Version]
  60. Ladarre, D.; Roland, A.B.; Biedzinski, S.; Ricobaraza, A.; Lenkei, Z. Polarized cellular patterns of endocannabinoid production and detection shape cannabinoid signaling in neurons. Front. Cell. Neurosci. 2014, 8, 426. [Google Scholar] [CrossRef] [Green Version]
  61. Leterrier, C.; Lainé, J.; Darmon, M.; Boudin, H.; Rossier, J.; Lenkei, Z. Constitutive activation drives compartment-selective endocytosis and axonal targeting of type 1 cannabinoid receptors. J. Neurosci. 2006, 26, 3141–3153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Nimchinsky, E.A.; Sabatini, B.L.; Svoboda, K. Structure and function of dendritic spines. Annu. Rev. Physiol. 2002, 64, 313–353. [Google Scholar] [CrossRef] [Green Version]
  63. Carvalho, A.F.; Reyes, B.A.S.; Ramalhosa, F.; Sousa, N.; Van Bockstaele, E.J. Repeated administration of a synthetic cannabinoid receptor agonist differentially affects cortical and accumbal neuronal morphology in adolescent and adult rats. Brain Struct. Funct. 2016, 221, 407–419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Spiga, S.; Lintas, A.; Diana, M. Altered Mesolimbic Dopamine System in THC Dependence. Curr. Neuropharmacol. 2011, 9, 200–204. [Google Scholar] [CrossRef] [Green Version]
  65. Spiga, S.; Lintas, A.; Migliore, M.; Diana, M. Altered architecture and functional consequences of the mesolimbic dopamine system in cannabis dependence. Addict. Biol. 2010, 15, 266–276. [Google Scholar] [CrossRef] [PubMed]
  66. Monday, H.R.; Bourdenx, M.; Jordan, B.A.; Castillo, P.E. Cb1-receptor-mediated inhibitory ltd triggers presynaptic remodeling via protein synthesis and ubiquitination. Elife 2020, 9, 1–25. [Google Scholar] [CrossRef] [PubMed]
  67. Incontro, S.; Sammari, M.; Azzaz, F.; Inglebert, Y.; Ankri, N.; Russier, M.; Fantini, J.; Debanne, D. Endocannabinoids tune intrinsic excitability in O-LM interneurons by direct modulation of post-synaptic Kv7 channels. J. Neurosci. 2021. [Google Scholar] [CrossRef]
  68. Okamura, K.; Tanaka, H.; Yagita, Y.; Saeki, Y.; Taguchi, A.; Hiraoka, Y.; Zeng, L.H.; Colman, D.R.; Miki, N. Cadherin activity is required for activity-induced spine remodeling. J. Cell Biol. 2004, 167, 961–972. [Google Scholar] [CrossRef]
  69. McKinney, R.A. Excitatory amino acid involvement in dendritic spine formation, maintenance and remodelling. J. Physiol. 2010, 588, 107–116. [Google Scholar] [CrossRef]
  70. Dani, J.A.; Bertrand, D. Nicotinic acetylcholine receptors and nicotinic cholinergic mechanisms of the central nervous system. Annu. Rev. Pharmacol. Toxicol. 2007, 47, 699–729. [Google Scholar] [CrossRef]
  71. Taly, A.; Corringer, P.-J.; Guedin, D.; Lestage, P.; Changeux, J.-P. Nicotinic receptors: Allosteric transitions and therapeutic targets in the nervous system. Nat. Rev. Drug Discov. 2009, 8, 733–750. [Google Scholar] [CrossRef] [PubMed]
  72. Livingstone, P.D.; Wonnacott, S. Nicotinic acetylcholine receptors and the ascending dopamine pathways. Biochem. Pharmacol. 2009, 78, 744–755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Albuquerque, E.X.; Pereira, E.F.R.; Alkondon, M.; Rogers, S.W. Mammalian nicotinic acetylcholine receptors: From structure to function. Physiol. Rev. 2009, 89, 73–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Gotti, C.; Clementi, F.; Fornari, A.; Gaimarri, A.; Guiducci, S.; Manfredi, I.; Moretti, M.; Pedrazzi, P.; Pucci, L.; Zoli, M. Structural and functional diversity of native brain neuronal nicotinic receptors. Biochem. Pharmacol. 2009, 78, 703–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. McQuiston, A.R. Acetylcholine release and inhibitory interneuron activity in hippocampal CA1. Front. Synaptic Neurosci. 2014, 6, 20. [Google Scholar] [CrossRef]
  76. Marchi, M.; Risso, F.; Viola, C.; Cavazzani, P.; Raiteri, M. Direct evidence that release-stimulating alpha7* nicotinic cholinergic receptors are localized on human and rat brain glutamatergic axon terminals. J. Neurochem. 2002, 80, 1071–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Lozada, A.F.; Wang, X.; Gounko, N.V.; Massey, K.A.; Duan, J.; Liu, Z.; Berg, D.K. Induction of dendritic spines by β2-containing nicotinic receptors. J. Neurosci. 2012, 32, 8391–8400. [Google Scholar] [CrossRef] [PubMed]
  78. Lozada, A.F.; Wang, X.; Gounko, N.V.; Massey, K.A.; Duan, J.; Liu, Z.; Berg, D.K. Glutamatergic synapse formation is promoted by α7-containing nicotinic acetylcholine receptors. J. Neurosci. 2012, 32, 7651–7661. [Google Scholar] [CrossRef] [Green Version]
  79. Halff, A.W.; Gómez-Varela, D.; John, D.; Berg, D.K. A Novel Mechanism for Nicotinic Potentiation of Glutamatergic Synapses. J. Neurosci. 2014, 34, 2051–2064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Balfour, D.J.K. The role of mesoaccumbens dopamine in nicotine dependence. Curr. Top. Behav. Neurosci. 2015, 24, 55–98. [Google Scholar] [CrossRef] [PubMed]
  81. Pidoplichko, V.I.; DeBiasi, M.; Williams, J.T.; Dani, J.A. Nicotine activates and desensitizes midbrain dopamine neurons. Nature 1997, 390, 401–404. [Google Scholar] [CrossRef]
  82. Corrigall, W.A.; Franklin, K.B.; Coen, K.M.; Clarke, P.B. The mesolimbic dopaminergic system is implicated in the reinforcing effects of nicotine. Psychopharmacology 1992, 107, 285–289. [Google Scholar] [CrossRef] [PubMed]
  83. Di Chiara, G. Role of dopamine in the behavioural actions of nicotine related to addiction. Eur. J. Pharmacol. 2000, 393, 295–314. [Google Scholar] [CrossRef]
  84. Wise, R.A. Dopamine, learning and motivation. Nat. Rev. Neurosci. 2004, 5, 483–494. [Google Scholar] [CrossRef]
  85. Watkins, S.S.; Epping-Jordan, M.P.; Koob, G.F.; Markou, A. Blockade of nicotine self-administration with nicotinic antagonists in rats. Pharmacol. Biochem. Behav. 1999, 62, 743–751. [Google Scholar] [CrossRef]
  86. Cohen, C.; Bergis, O.E.; Galli, F.; Lochead, A.W.; Jegham, S.; Biton, B.; Leonardon, J.; Avenet, P.; Sgard, F.; Besnard, F.; et al. SSR591813, a novel selective and partial alpha4beta2 nicotinic receptor agonist with potential as an aid to smoking cessation. J. Pharmacol. Exp. Ther. 2003, 306, 407–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Picciotto, M.R.; Zoli, M.; Rimondini, R.; Léna, C.; Marubio, L.M.; Pich, E.M.; Fuxe, K.; Changeux, J.P. Acetylcholine receptors containing the beta2 subunit are involved in the reinforcing properties of nicotine. Nature 1998, 391, 173–177. [Google Scholar] [CrossRef]
  88. Marubio, L.M.; Gardier, A.M.; Durier, S.; David, D.; Klink, R.; Arroyo-Jimenez, M.M.; McIntosh, J.M.; Rossi, F.; Champtiaux, N.; Zoli, M.; et al. Effects of nicotine in the dopaminergic system of mice lacking the alpha4 subunit of neuronal nicotinic acetylcholine receptors. Eur. J. Neurosci. 2003, 17, 1329–1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Xi, Z.; Spiller, K.; Gardner, E.L. Mechanism-based medication development for the treatment of nicotine dependence. Acta Pharmacol. Sin. 2009, 30, 723–739. [Google Scholar] [CrossRef] [Green Version]
  90. Scott, L.; Zelenin, S.; Malmersjö, S.; Kowalewski, J.M.; Markus, E.Z.; Nairn, A.C.; Greengard, P.; Brismar, H.; Aperia, A. Allosteric changes of the NMDA receptor trap diffusible dopamine 1 receptors in spines. Proc. Natl. Acad. Sci. USA 2006, 103, 762–767. [Google Scholar] [CrossRef] [Green Version]
  91. Cepeda, C.; Levine, M.S. Where do you think you are going? The NMDA-D1 receptor trap. Sci. STKE 2006, 2006, pe20. [Google Scholar] [CrossRef]
  92. Calabresi, P.; Gubellini, P.; Centonze, D.; Picconi, B.; Bernardi, G.; Chergui, K.; Svenningsson, P.; Fienberg, A.A.; Greengard, P. Dopamine and cAMP-regulated phosphoprotein 32 kDa controls both striatal long-term depression and long-term potentiation, opposing forms of synaptic plasticity. J. Neurosci. 2000, 20, 8443–8451. [Google Scholar] [CrossRef]
  93. Kerr, J.N.; Wickens, J.R. Dopamine D-1/D-5 receptor activation is required for long-term potentiation in the rat neostriatum in vitro. J. Neurophysiol. 2001, 85, 117–124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Sawaguchi, T.; Goldman-Rakic, P.S. The role of D1-dopamine receptor in working memory: Local injections of dopamine antagonists into the prefrontal cortex of rhesus monkeys performing an oculomotor delayed-response task. J. Neurophysiol. 1994, 71, 515–528. [Google Scholar] [CrossRef] [PubMed]
  95. Fiorentini, C.; Gardoni, F.; Spano, P.; Di Luca, M.; Missale, C. Regulation of dopamine D1 receptor trafficking and desensitization by oligomerization with glutamate N-methyl-D-aspartate receptors. J. Biol. Chem. 2003, 278, 20196–20202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Maurer, S.V.; Williams, C.L. The Cholinergic System Modulates Memory and Hippocampal Plasticity via Its Interactions with Non-Neuronal Cells. Front. Immunol. 2017, 8, 1489. [Google Scholar] [CrossRef] [Green Version]
  97. Prado, V.F.; Janickova, H.; Al-Onaizi, M.A.; Prado, M.A.M. Cholinergic circuits in cognitive flexibility. Neuroscience 2017, 345, 130–141. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, L.; Almeida, L.E.F.; Spornick, N.A.; Kenyon, N.; Kamimura, S.; Khaibullina, A.; Nouraie, M.; Quezado, Z.M.N. Modulation of social deficits and repetitive behaviors in a mouse model of autism: The role of the nicotinic cholinergic system. Psychopharmacology 2015, 232, 4303–4316. [Google Scholar] [CrossRef] [PubMed]
  99. Oz, M. Receptor-independent actions of cannabinoids on cell membranes: Focus on endocannabinoids. Pharmacol. Ther. 2006, 111, 114–144. [Google Scholar] [CrossRef] [PubMed]
  100. Pertwee, R.G. Receptors and channels targeted by synthetic cannabinoid receptor agonists and antagonists. Curr. Med. Chem. 2010, 17, 1360–1381. [Google Scholar] [CrossRef] [Green Version]
  101. Scherma, M.; Muntoni, A.; Melis, M.; Fattore, L.; Fadda, P.; Fratta, W.; Pistis, M. Interactions between the endocannabinoid and nicotinic cholinergic systems: Preclinical evidence and therapeutic perspectives. Psychopharmacology 2016, 233, 1765–1777. [Google Scholar] [CrossRef]
  102. Jackson, K.J.; Marks, M.J.; Vann, R.E.; Chen, X.; Gamage, T.F.; Warner, J.A.; Damaj, M.I. Role of alpha5 nicotinic acetylcholine receptors in pharmacological and behavioral effects of nicotine in mice. J. Pharmacol. Exp. Ther. 2010, 334, 137–146. [Google Scholar] [CrossRef] [Green Version]
  103. Howlett, A.C.; Bidaut-Russell, M.; Devane, W.A.; Melvin, L.S.; Johnson, M.R.; Herkenham, M. The cannabinoid receptor: Biochemical, anatomical and behavioral characterization. Trends Neurosci. 1990, 13, 420–423. [Google Scholar] [CrossRef]
  104. Lichtman, A.H.; Cook, S.A.; Martin, B.R. Investigation of brain sites mediating cannabinoid-induced antinociception in rats: Evidence supporting periaqueductal gray involvement. J. Pharmacol. Exp. Ther. 1996, 276, 585–593. [Google Scholar] [PubMed]
  105. Sañudo-Peña, M.C.; Romero, J.; Seale, G.E.; Fernandez-Ruiz, J.J.; Walker, J.M. Activational role of cannabinoids on movement. Eur. J. Pharmacol. 2000, 391, 269–274. [Google Scholar] [CrossRef]
  106. Justinova, Z.; Goldberg, S.R.; Heishman, S.J.; Tanda, G. Self-administration of cannabinoids by experimental animals and human marijuana smokers. Pharmacol. Biochem. Behav. 2005, 81, 285–299. [Google Scholar] [CrossRef] [Green Version]
  107. Ahsan, H.M.; de la Peña, J.B.I.; Botanas, C.J.; Kim, H.J.; Yu, G.Y.; Cheong, J.H. Conditioned place preference and self-administration induced by nicotine in adolescent and adult rats. Biomol. Ther. 2014, 22, 460–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Scherma, M.; Justinová, Z.; Zanettini, C.; Panlilio, L.V.; Mascia, P.; Fadda, P.; Fratta, W.; Makriyannis, A.; Vadivel, S.K.; Gamaleddin, I.; et al. The anandamide transport inhibitor AM404 reduces the rewarding effects of nicotine and nicotine-induced dopamine elevations in the nucleus accumbens shell in rats. Br. J. Pharmacol. 2012, 165, 2539–2548. [Google Scholar] [CrossRef] [Green Version]
  109. Balerio, G.N.; Aso, E.; Berrendero, F.; Murtra, P.; Maldonado, R. Delta9-tetrahydrocannabinol decreases somatic and motivational manifestations of nicotine withdrawal in mice. Eur. J. Neurosci. 2004, 20, 2737–2748. [Google Scholar] [CrossRef]
  110. Balerio, G.N.; Aso, E.; Maldonado, R. Role of the cannabinoid system in the effects induced by nicotine on anxiety-like behaviour in mice. Psychopharmacology 2006, 184, 504–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Valjent, E.; Mitchell, J.M.; Besson, M.-J.; Caboche, J.; Maldonado, R. Behavioural and biochemical evidence for interactions between Delta 9-tetrahydrocannabinol and nicotine. Br. J. Pharmacol. 2002, 135, 564–578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Acquas, E.; Pisanu, A.; Marrocu, P.; Goldberg, S.R.; Di Chiara, G. Δ9-tetrahydrocannabinol enhances cortical and hippocampal acetylcholine release in vivo: A microdialysis study. Eur. J. Pharmacol. 2001, 419, 155–161. [Google Scholar] [CrossRef]
  113. Pisanu, A.; Acquas, E.; Fenu, S.; Di Chiara, G. Modulation of Δ9-THC-induced increase of cortical and hippocampal acetylcholine release by μ opioid and D1 dopamine receptors. Neuropharmacology 2006, 50, 661–670. [Google Scholar] [CrossRef] [PubMed]
  114. Tzavara, E.T.; Wade, M.; Nomikos, G.G. Biphasic effects of cannabinoids on acetylcholine release in the hippocampus: Site and mechanism of action. J. Neurosci. 2003, 23, 9374–9384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Gamaleddin, I.H.; Trigo, J.M.; Gueye, A.B.; Zvonok, A.; Makriyannis, A.; Goldberg, S.R.; Le Foll, B. Role of the endogenous cannabinoid system in nicotine addiction: Novel insights. Front. Psychiatry 2015, 6, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Solinas, M.; Scherma, M.; Tanda, G.; Wertheim, C.E.; Fratta, W.; Goldberg, S.R. Nicotinic facilitation of delta9-tetrahydrocannabinol discrimination involves endogenous anandamide. J. Pharmacol. Exp. Ther. 2007, 321, 1127–1134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Oz, M.; Ravindran, A.; Diaz-Ruiz, O.; Zhang, L.; Morales, M. The endogenous cannabinoid anandamide inhibits alpha7 nicotinic acetylcholine receptor-mediated responses in Xenopus oocytes. J. Pharmacol. Exp. Ther. 2003, 306, 1003–1010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Oz, M.; Zhang, L.; Ravindran, A.; Morales, M.; Lupica, C.R. Differential effects of endogenous and synthetic cannabinoids on alpha7-nicotinic acetylcholine receptor-mediated responses in Xenopus Oocytes. J. Pharmacol. Exp. Ther. 2004, 310, 1152–1160. [Google Scholar] [CrossRef] [Green Version]
  119. Lagalwar, S.; Bordayo, E.Z.; Hoffmann, K.L.; Fawcett, J.R.; Frey, W.H. 2nd Anandamides inhibit binding to the muscarinic acetylcholine receptor. J. Mol. Neurosci. 1999, 13, 55–61. [Google Scholar] [CrossRef]
  120. Spivak, C.E.; Lupica, C.R.; Oz, M. The Endocannabinoid Anandamide Inhibits the Function of α4β2 Nicotinic Acetylcholine Receptors. Mol. Pharmacol. 2007, 72, 1024–1032. [Google Scholar] [CrossRef] [Green Version]
  121. Oz, M. Receptor-independent effects of endocannabinoids on ion channels. Curr. Pharm. Des. 2006, 12, 227–239. [Google Scholar] [CrossRef]
  122. Oz, M.; Jackson, S.N.; Woods, A.S.; Morales, M.; Zhang, L. Additive effects of endogenous cannabinoid anandamide and ethanol on alpha7-nicotinic acetylcholine receptor-mediated responses in Xenopus Oocytes. J. Pharmacol. Exp. Ther. 2005, 313, 1272–1280. [Google Scholar] [CrossRef] [PubMed]
  123. Jackson, S.N.; Singhal, S.K.; Woods, A.S.; Morales, M.; Shippenberg, T.; Zhang, L.; Oz, M. Volatile anesthetics and endogenous cannabinoid anandamide have additive and independent inhibitory effects on alpha(7)-nicotinic acetylcholine receptor-mediated responses in Xenopus oocytes. Eur. J. Pharmacol. 2008, 582, 42–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Sabatini, B.L.; Oertner, T.G.; Svoboda, K. The life cycle of Ca(2+) ions in dendritic spines. Neuron 2002, 33, 439–452. [Google Scholar] [CrossRef] [Green Version]
  125. Fujii, H.; Inoue, M.; Okuno, H.; Sano, Y.; Takemoto-Kimura, S.; Kitamura, K.; Kano, M.; Bito, H. Nonlinear decoding and asymmetric representation of neuronal input information by CaMKIIα and calcineurin. Cell Rep. 2013, 3, 978–987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Chang, J.-Y.; Parra-Bueno, P.; Laviv, T.; Szatmari, E.M.; Lee, S.-J.R.; Yasuda, R. CaMKII Autophosphorylation Is Necessary for Optimal Integration of Ca(2+) Signals during LTP Induction, but Not Maintenance. Neuron 2017, 94, 800–808. [Google Scholar] [CrossRef] [Green Version]
  127. Murakoshi, H.; Wang, H.; Yasuda, R. Local, persistent activation of Rho GTPases during plasticity of single dendritic spines. Nature 2011, 472, 100. [Google Scholar] [CrossRef] [PubMed]
  128. Borovac, J.; Bosch, M.; Okamoto, K. Regulation of actin dynamics during structural plasticity of dendritic spines: Signaling messengers and actin-binding proteins. Mol. Cell. Neurosci. 2018, 91, 122–130. [Google Scholar] [CrossRef]
  129. Costa, J.F.; Dines, M.; Lamprecht, R. The Role of Rac GTPase in Dendritic Spine Morphogenesis and Memory. Front. Synaptic Neurosci. 2020, 12, 12. [Google Scholar] [CrossRef]
  130. Chernomordik, L.V.; Kozlov, M.M. Protein-lipid interplay in fusion and fission of biological membranes. Annu. Rev. Biochem. 2003, 72, 175–207. [Google Scholar] [CrossRef]
  131. Puchkov, D.; Haucke, V. Greasing the synaptic vesicle cycle by membrane lipids. Trends Cell Biol. 2013, 23, 493–503. [Google Scholar] [CrossRef] [PubMed]
  132. Dieterich, D.C.; Kreutz, M.R. Proteomics of the Synapse—A Quantitative Approach to Neuronal Plasticity *. Mol. Cell. Proteomics 2016, 15, 368–381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Jurado, S.; Benoist, M.; Lario, A.; Knafo, S.; Petrok, C.N.; Esteban, J.A. PTEN is recruited to the postsynaptic terminal for NMDA receptor-dependent long-term depression. EMBO J. 2010, 29, 2827–2840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Martin, M.G.; Ahmed, T.; Korovaichuk, A.; Venero, C.; Menchón, S.A.; Salas, I.; Munck, S.; Herreras, O.; Balschun, D.; Dotti, C.G. Constitutive hippocampal cholesterol loss underlies poor cognition in old rodents. EMBO Mol. Med. 2014, 6, 902–917. [Google Scholar] [CrossRef]
  135. Tulodziecka, K.; Diaz-Rohrer, B.B.; Farley, M.M.; Chan, R.B.; Di Paolo, G.; Levental, K.R.; Waxham, M.N.; Levental, I. Remodeling of the postsynaptic plasma membrane during neural development. Mol. Biol. Cell 2016, 27, 3480–3489. [Google Scholar] [CrossRef]
  136. Merrill, A.H.J. Sphingolipid and glycosphingolipid metabolic pathways in the era of sphingolipidomics. Chem. Rev. 2011, 111, 6387–6422. [Google Scholar] [CrossRef]
  137. Borgmeyer, M.; Coman, C.; Has, C.; Schött, H.-F.; Li, T.; Westhoff, P.; Cheung, Y.F.H.; Hoffmann, N.; Yuanxiang, P.; Behnisch, T.; et al. Multiomics of synaptic junctions reveals altered lipid metabolism and signaling following environmental enrichment. Cell Rep. 2021, 37, 109797. [Google Scholar] [CrossRef]
  138. van Meer, G.; Voelker, D.R.; Feigenson, G.W. Membrane lipids: Where they are and how they behave. Nat. Rev. Mol. Cell Biol. 2008, 9, 112–124. [Google Scholar] [CrossRef]
  139. Casares, D.; Escribá, P.V.; Rosselló, C.A. Membrane Lipid Composition: Effect on Membrane and Organelle Structure, Function and Compartmentalization and Therapeutic Avenues. Int. J. Mol. Sci. 2019, 20, 2167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Goñi, F.M.; Sot, J.; Alonso, A. Biophysical properties of sphingosine, ceramides and other simple sphingolipids. Biochem. Soc. Trans. 2014, 42, 1401–1408. [Google Scholar] [CrossRef]
  141. Sohn, J.; Lin, H.; Fritch, M.R.; Tuan, R.S. Influence of cholesterol/caveolin-1/caveolae homeostasis on membrane properties and substrate adhesion characteristics of adult human mesenchymal stem cells. Stem Cell Res. Ther. 2018, 9, 86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Song, E.S.; Oh, Y.; Sung, B.J. Interdomain exchange and the flip-flop of cholesterol in ternary component lipid membranes and their effects on heterogeneous cholesterol diffusion. Phys. Rev. E 2021, 104, 44402. [Google Scholar] [CrossRef]
  143. Baier, C.J.; Fantini, J.; Barrantes, F.J. Disclosure of cholesterol recognition motifs in transmembrane domains of the human nicotinic acetylcholine receptor. Sci. Rep. 2011, 1, 69. [Google Scholar] [CrossRef] [Green Version]
  144. Fantini, J.; Di Scala, C.; Evans, L.S.; Williamson, P.T.F.; Barrantes, F.J. A mirror code for protein-cholesterol interactions in the two leaflets of biological membranes. Sci. Rep. 2016, 6, 21907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Oddi, S.; Dainese, E.; Fezza, F.; Lanuti, M.; Barcaroli, D.; De Laurenzi, V.; Centonze, D.; Maccarrone, M. Functional characterization of putative cholesterol binding sequence (CRAC) in human type-1 cannabinoid receptor. J. Neurochem. 2011, 116, 858–865. [Google Scholar] [CrossRef]
  146. Hua, T.; Vemuri, K.; Pu, M.; Qu, L.; Han, G.W.; Wu, Y.; Zhao, S.; Shui, W.; Li, S.; Korde, A.; et al. Crystal Structure of the Human Cannabinoid Receptor CB(1). Cell 2016, 167, 750–762. [Google Scholar] [CrossRef] [Green Version]
  147. Krishna Kumar, K.; Shalev-Benami, M.; Robertson, M.J.; Hu, H.; Banister, S.D.; Hollingsworth, S.A.; Latorraca, N.R.; Kato, H.E.; Hilger, D.; Maeda, S.; et al. Structure of a Signaling Cannabinoid Receptor 1-G Protein Complex. Cell 2019, 176, 448–458. [Google Scholar] [CrossRef] [Green Version]
  148. Allende, M.L.; Zhu, H.; Kono, M.; Hoachlander-Hobby, L.E.; Huso, V.L.; Proia, R.L. Genetic defects in the sphingolipid degradation pathway and their effects on microglia in neurodegenerative disease. Cell. Signal. 2021, 78, 109879. [Google Scholar] [CrossRef]
  149. Hering, H.; Lin, C.C.; Sheng, M. Lipid rafts in the maintenance of synapses, dendritic spines, and surface AMPA receptor stability. J. Neurosci. 2003, 23, 3262–3271. [Google Scholar] [CrossRef] [Green Version]
  150. Carrasco, P.; Sahún, I.; McDonald, J.; Ramírez, S.; Jacas, J.; Gratacós, E.; Sierra, A.Y.; Serra, D.; Herrero, L.; Acker-Palmer, A.; et al. Ceramide levels regulated by carnitine palmitoyltransferase 1C control dendritic spine maturation and cognition. J. Biol. Chem. 2012, 287, 21224–21232. [Google Scholar] [CrossRef] [Green Version]
  151. Hammond, G.R.V.; Schiavo, G. Polyphosphoinositol lipids: Under-PPInning synaptic function in health and disease. Dev. Neurobiol. 2007, 67, 1232–1247. [Google Scholar] [CrossRef] [PubMed]
  152. Richards, D.A.; Mateos, J.M.; Hugel, S.; De Paola, V.; Caroni, P.; Gähwiler, B.H.; McKinney, R.A. Glutamate induces the rapid formation of spine head protrusions in hippocampal slice cultures. Proc. Natl. Acad. Sci. USA 2005, 102, 6166–6171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Ueda, Y.; Hayashi, Y. PIP3 regulates Spinule formation in dendritic spines during structural long-term potentiation. J. Neurosci. 2013, 33, 11040–11047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Yin, H.L.; Janmey, P.A. Phosphoinositide Regulation of the Actin Cytoskeleton. Annu. Rev. Physiol. 2003, 65, 761–789. [Google Scholar] [CrossRef]
  155. Kumar, V.; Zhang, M.X.; Swank, M.W.; Kunz, J.; Wu, G.Y. Regulation of dendritic morphogenesis by Ras-PI3K-Akt-mTOR and Ras-MAPK signaling pathways. J. Neurosci. 2005, 25, 11288–11299. [Google Scholar] [CrossRef] [Green Version]
  156. Kelleher, R.J., 3rd; Govindarajan, A.; Tonegawa, S. Translational regulatory mechanisms in persistent forms of synaptic plasticity. Neuron 2004, 44, 59–73. [Google Scholar] [CrossRef] [Green Version]
  157. McLaughlin, S.; Murray, D. Plasma membrane phosphoinositide organization by protein electrostatics. Nature 2005, 438, 605–611. [Google Scholar] [CrossRef]
  158. Calabrese, B.; Halpain, S. Essential role for the PKC target MARCKS in maintaining dendritic spine morphology. Neuron 2005, 48, 77–90. [Google Scholar] [CrossRef] [Green Version]
  159. Horne, E.A.; Dell’Acqua, M.L. Phospholipase C is required for changes in postsynaptic structure and function associated with NMDA receptor-dependent long-term depression. J. Neurosci. 2007, 27, 3523–3534. [Google Scholar] [CrossRef]
  160. Arroyo, A.I.; Camoletto, P.G.; Morando, L.; Sassoe-Pognetto, M.; Giustetto, M.; Van Veldhoven, P.P.; Schuchman, E.H.; Ledesma, M.D. Pharmacological reversion of sphingomyelin-induced dendritic spine anomalies in a Niemann Pick disease type A mouse model. EMBO Mol. Med. 2014, 6, 398–413. [Google Scholar] [CrossRef] [Green Version]
  161. Stoffel, W. Functional analysis of acid and neutral sphingomyelinases in vitro and in vivo. Chem. Phys. Lipids 1999, 102, 107–121. [Google Scholar] [CrossRef]
  162. Franco-Villanueva, A.; Fernández-López, E.; Gabandé-Rodríguez, E.; Bañón-Rodríguez, I.; Esteban, J.A.; Antón, I.M.; Ledesma, M.D. WIP modulates dendritic spine actin cytoskeleton by transcriptional control of lipid metabolic enzymes. Hum. Mol. Genet. 2014, 23, 4383–4395. [Google Scholar] [CrossRef] [Green Version]
  163. Barrantes, F.J. Structural basis for lipid modulation of nicotinic acetylcholine receptor function. Brain Res. Brain Res. Rev. 2004, 47, 71–95. [Google Scholar] [CrossRef]
  164. Barrantes, F.J. Cholesterol effects on nicotinic acetylcholine receptor. J. Neurochem. 2007, 103, 72–80. [Google Scholar] [CrossRef] [PubMed]
  165. Baier, C.J.; Gallegos, C.E.; Levi, V.; Barrantes, F.J. Cholesterol modulation of nicotinic acetylcholine receptor surface mobility. Eur. Biophys. J. 2010, 39, 213–227. [Google Scholar] [CrossRef]
  166. Mosqueira, A.; Camino, P.A.; Barrantes, F.J. Antibody-induced crosslinking and cholesterol-sensitive, anomalous diffusion of nicotinic acetylcholine receptors. J. Neurochem. 2020, 152, 663–674. [Google Scholar] [CrossRef]
  167. Mosqueira, A.; Camino, P.A.; Barrantes, F.J. Cholesterol modulates acetylcholine receptor diffusion by tuning confinement sojourns and nanocluster stability. Sci. Rep. 2018, 8, 11974. [Google Scholar] [CrossRef]
  168. Pediconi, M.F.; Gallegos, C.E.; De Los Santos, E.B.; Barrantes, F.J. Metabolic cholesterol depletion hinders cell-surface trafficking of the nicotinic acetylcholine receptor. Neuroscience 2004, 128, 239–249. [Google Scholar] [CrossRef] [PubMed]
  169. Borroni, V.; Kamerbeek, C.; Pediconi, M.F.; Barrantes, F.J. Lovastatin Differentially Regulates α7 and α4 Neuronal Nicotinic Acetylcholine Receptor Levels in Rat Hippocampal Neurons. Molecules 2020, 25, 4838. [Google Scholar] [CrossRef] [PubMed]
  170. Baier, C.J.; Barrantes, F.J. Sphingolipids are necessary for nicotinic acetylcholine receptor export in the early secretory pathway. J. Neurochem. 2007, 101, 1072–1084. [Google Scholar] [CrossRef]
  171. Yeliseev, A.; Iyer, M.R.; Joseph, T.T.; Coffey, N.J.; Cinar, R.; Zoubak, L.; Kunos, G.; Gawrisch, K. Cholesterol as a modulator of cannabinoid receptor CB2 signaling. Sci. Reports 2021 111 2021, 11, 1–18. [Google Scholar] [CrossRef] [PubMed]
  172. Bari, M.; Spagnuolo, P.; Fezza, F.; Oddi, S.; Pasquariello, N.; Finazzi-Agrò, A.; Maccarrone, M. Effect of lipid rafts on Cb2 receptor signaling and 2-arachidonoyl-glycerol metabolism in human immune cells. J. Immunol. 2006, 177, 4971–4980. [Google Scholar] [CrossRef] [Green Version]
  173. Di Scala, C.; Fantini, J.; Yahi, N.; Barrantes, F.J.; Chahinian, H. Anandamide Revisited: How Cholesterol and Ceramides Control Receptor-Dependent and Receptor-Independent Signal Transmission Pathways of a Lipid Neurotransmitter. Biomolecules 2018, 8, 31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Kapus, A.; Janmey, P. Plasma membrane--cortical cytoskeleton interactions: A cell biology approach with biophysical considerations. Compr. Physiol. 2013, 3, 1231–1281. [Google Scholar] [CrossRef]
  175. Kusumi, A.; Nakada, C.; Ritchie, K.; Murase, K.; Suzuki, K.; Murakoshi, H.; Kasai, R.S.; Kondo, J.; Fujiwara, T. Paradigm shift of the plasma membrane concept from the two-dimensional continuum fluid to the partitioned fluid: High-speed single-molecule tracking of membrane molecules. Annu. Rev. Biophys. Biomol. Struct. 2005, 34, 351–378. [Google Scholar] [CrossRef] [Green Version]
  176. Andrade, D.M.; Clausen, M.P.; Keller, J.; Mueller, V.; Wu, C.; Bear, J.E.; Hell, S.W.; Lagerholm, B.C.; Eggeling, C. Cortical actin networks induce spatio-temporal confinement of phospholipids in the plasma membrane--a minimally invasive investigation by STED-FCS. Sci. Rep. 2015, 5, 11454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Fujiwara, T.; Ritchie, K.; Murakoshi, H.; Jacobson, K.; Kusumi, A. Phospholipids undergo hop diffusion in compartmentalized cell membrane. J. Cell Biol. 2002, 157, 1071–1081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Auth, T.; Gov, N.S. Diffusion in a fluid membrane with a flexible cortical cytoskeleton. Biophys. J. 2009, 96, 818–830. [Google Scholar] [CrossRef] [Green Version]
  179. Hotulainen, P.; Hoogenraad, C.C. Actin in dendritic spines: Connecting dynamics to function. J. Cell Biol. 2010, 189, 619–629. [Google Scholar] [CrossRef] [Green Version]
  180. Kwik, J.; Boyle, S.; Fooksman, D.; Margolis, L.; Sheetz, M.P.; Edidin, M. Membrane cholesterol, lateral mobility, and the phosphatidylinositol 4,5-bisphosphate-dependent organization of cell actin. Proc. Natl. Acad. Sci. USA 2003, 100, 13964–13969. [Google Scholar] [CrossRef] [Green Version]
  181. Cingolani, L.A.; Goda, Y. Actin in action: The interplay between the actin cytoskeleton and synaptic efficacy. Nat. Rev. Neurosci. 2008, 9, 344–356. [Google Scholar] [CrossRef] [PubMed]
  182. Nelson, J.C.; Stavoe, A.K.H.; Colón-Ramos, D.A. The actin cytoskeleton in presynaptic assembly. Cell Adh. Migr. 2013, 7, 379–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Michel, K.; Müller, J.A.; Oprişoreanu, A.-M.; Schoch, S. The presynaptic active zone: A dynamic scaffold that regulates synaptic efficacy. Exp. Cell Res. 2015, 335, 157–164. [Google Scholar] [CrossRef] [PubMed]
  184. Rust, M.B.; Maritzen, T. Relevance of presynaptic actin dynamics for synapse function and mouse behavior. Exp. Cell Res. 2015, 335, 165–171. [Google Scholar] [CrossRef]
  185. Bartol, T.M.; Bromer, C.; Kinney, J.; Chirillo, M.A.; Bourne, J.N.; Harris, K.M.; Sejnowski, T.J. Nanoconnectomic upper bound on the variability of synaptic plasticity. Elife 2015, 4, e10778. [Google Scholar] [CrossRef] [Green Version]
  186. Bourne, J.N.; Chirillo, M.A.; Harris, K.M. Presynaptic ultrastructural plasticity along CA3→CA1 axons during long-term potentiation in mature hippocampus. J. Comp. Neurol. 2013, 521, 3898–3912. [Google Scholar] [CrossRef] [Green Version]
  187. Meyer, D.; Bonhoeffer, T.; Scheuss, V. Balance and stability of synaptic structures during synaptic plasticity. Neuron 2014, 82, 430–443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Gundelfinger, E.D.; Fejtova, A. Molecular organization and plasticity of the cytomatrix at the active zone. Curr. Opin. Neurobiol. 2012, 22, 423–430. [Google Scholar] [CrossRef]
  189. Matz, J.; Gilyan, A.; Kolar, A.; McCarvill, T.; Krueger, S.R. Rapid structural alterations of the active zone lead to sustained changes in neurotransmitter release. Proc. Natl. Acad. Sci. USA 2010, 107, 8836–8841. [Google Scholar] [CrossRef] [Green Version]
  190. Monday, H.R.; Castillo, P.E. Closing the gap: Long-term presynaptic plasticity in brain function and disease. Curr. Opin. Neurobiol. 2017, 45, 106–112. [Google Scholar] [CrossRef] [PubMed]
  191. Runge, K.; Cardoso, C.; de Chevigny, A. Dendritic Spine Plasticity: Function and Mechanisms. Front. Synaptic Neurosci. 2020, 12, 36. [Google Scholar] [CrossRef] [PubMed]
  192. Nakahata, Y.; Yasuda, R. Plasticity of Spine Structure: Local Signaling, Translation and Cytoskeletal Reorganization. Front. Synaptic Neurosci. 2018, 10, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Holcman, D.; Triller, A. Modeling synaptic dynamics driven by receptor lateral diffusion. Biophys. J. 2006, 91, 2405–2415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Yang, X.; Specht, C.G. Subsynaptic Domains in Super-Resolution Microscopy: The Treachery of Images. Front. Mol. Neurosci. 2019, 12, 161. [Google Scholar] [CrossRef]
  195. Hruska, M.; Henderson, N.; Le Marchand, S.J.; Jafri, H.; Dalva, M.B. Synaptic nanomodules underlie the organization and plasticity of spine synapses. Nat. Neurosci. 2018, 21, 671–682. [Google Scholar] [CrossRef]
  196. Wegner, W.; Mott, A.C.; Grant, S.G.N.; Steffens, H.; Willig, K.I. In vivo STED microscopy visualizes PSD95 sub-structures and morphological changes over several hours in the mouse visual cortex. Sci. Rep. 2018, 8, 219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Moretto, E.; Longatti, A.; Murru, L.; Chamma, I.; Sessa, A.; Zapata, J.; Hosy, E.; Sainlos, M.; Saint-Pol, J.; Rubinstein, E.; et al. TSPAN5 Enriched Microdomains Provide a Platform for Dendritic Spine Maturation through Neuroligin-1 Clustering. Cell Rep. 2019, 29, 1130–1146. [Google Scholar] [CrossRef] [Green Version]
  198. Yang, X.; Annaert, W. The Nanoscopic Organization of Synapse Structures: A Common Basis for Cell Communication. Membranes 2021, 11, 248. [Google Scholar] [CrossRef]
  199. Triller, A.; Choquet, D. Surface trafficking of receptors between synaptic and extrasynaptic membranes: And yet they do move! Trends Neurosci. 2005, 28, 133–139. [Google Scholar] [CrossRef]
  200. Song, I.; Huganir, R.L. Regulation of AMPA receptors during synaptic plasticity. Trends Neurosci. 2002, 25, 578–588. [Google Scholar] [CrossRef]
  201. Choquet, D.; Triller, A. The role of receptor diffusion in the organization of the postsynaptic membrane. Nat. Rev. Neurosci. 2003, 4, 251–265. [Google Scholar] [CrossRef]
  202. Triller, A.; Choquet, D. New concepts in synaptic biology derived from single-molecule imaging. Neuron 2008, 59, 359–374. [Google Scholar] [CrossRef] [Green Version]
  203. Blanpied, T.A.; Scott, D.B.; Ehlers, M.D. Dynamics and regulation of clathrin coats at specialized endocytic zones of dendrites and spines. Neuron 2002, 36, 435–449. [Google Scholar] [CrossRef] [Green Version]
  204. Rácz, B.; Blanpied, T.A.; Ehlers, M.D.; Weinberg, R.J. Lateral organization of endocytic machinery in dendritic spines. Nat. Neurosci. 2004, 7, 917–918. [Google Scholar] [CrossRef]
  205. Lu, J.; Helton, T.D.; Blanpied, T.A.; Rácz, B.; Newpher, T.M.; Weinberg, R.J.; Ehlers, M.D. Postsynaptic positioning of endocytic zones and AMPA receptor cycling by physical coupling of dynamin-3 to Homer. Neuron 2007, 55, 874–889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Groc, L.; Choquet, D. Measurement and characteristics of neurotransmitter receptor surface trafficking (Review). Mol. Membr. Biol. 2008, 25, 344–352. [Google Scholar] [CrossRef] [PubMed]
  207. Kusters, R.; van der Heijden, T.; Kaoui, B.; Harting, J.; Storm, C. Forced transport of deformable containers through narrow constrictions. Phys. Rev. E. Stat. Nonlin. Soft Matter Phys. 2014, 90, 33006. [Google Scholar] [CrossRef] [Green Version]
  208. Wang, J.; Fourriere, L.; Gleeson, P.A. Local Secretory Trafficking Pathways in Neurons and the Role of Dendritic Golgi Outposts in Different Cell Models. Front. Mol. Neurosci. 2020, 13, 597391. [Google Scholar] [CrossRef]
  209. Frotscher, M.; Studer, D.; Graber, W.; Chai, X.; Nestel, S.; Zhao, S. Fine structure of synapses on dendritic spines. Front. Neuroanat. 2014, 8, 94. [Google Scholar] [CrossRef] [Green Version]
  210. Renner, M.; Specht, C.G.; Triller, A. Molecular dynamics of postsynaptic receptors and scaffold proteins. Curr. Opin. Neurobiol. 2008, 18, 532–540. [Google Scholar] [CrossRef]
  211. Newpher, T.M.; Ehlers, M.D. Glutamate receptor dynamics in dendritic microdomains. Neuron 2008, 58, 472–497. [Google Scholar] [CrossRef] [Green Version]
  212. Bürli, T.; Baer, K.; Ewers, H.; Sidler, C.; Fuhrer, C.; Fritschy, J.-M. Single Particle Tracking of α7 Nicotinic AChR in Hippocampal Neurons Reveals Regulated Confinement at Glutamatergic and GABAergic Perisynaptic Sites. PLoS ONE 2010, 5, e11507. [Google Scholar] [CrossRef] [PubMed]
  213. Lee, S.-H.; Földy, C.; Soltesz, I. Distinct endocannabinoid control of GABA release at perisomatic and dendritic synapses in the hippocampus. J. Neurosci. 2010, 30, 7993–8000. [Google Scholar] [CrossRef]
  214. Dudok, B.; Barna, L.; Ledri, M.; Szabó, S.I.; Szabadits, E.; Pintér, B.; Woodhams, S.G.; Henstridge, C.M.; Balla, G.Y.; Nyilas, R.; et al. Cell-specific STORM superresolution imaging reveals nanoscale organization of cannabinoid signaling. Nat. Neurosci. 2015, 18, 75. [Google Scholar] [CrossRef]
  215. Goncalves, J.; Bartol, T.M.; Camus, C.; Levet, F.; Menegolla, A.P.; Sejnowski, T.J.; Sibarita, J.-B.; Vivaudou, M.; Choquet, D.; Hosy, E. Nanoscale co-organization and coactivation of AMPAR, NMDAR, and mGluR at excitatory synapses. Proc. Natl. Acad. Sci. USA 2020, 117, 14503–14511. [Google Scholar] [CrossRef]
  216. Cognet, L.; Groc, L.; Lounis, B.; Choquet, D. Multiple routes for glutamate receptor trafficking: Surface diffusion and membrane traffic cooperate to bring receptors to synapses. Sci. STKE 2006, 2006, pe13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Martini, L.; Waldhoer, M.; Pusch, M.; Kharazia, V.; Fong, J.; Lee, J.H.; Freissmuth, C.; Whistler, J.L. Ligand-induced down-regulation of the cannabinoid 1 receptor is mediated by the G-protein-coupled receptor-associated sorting protein GASP1. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2007, 21, 802–811. [Google Scholar] [CrossRef] [PubMed]
  218. Jin, W.; Brown, S.; Roche, J.P.; Hsieh, C.; Celver, J.P.; Kovoor, A.; Chavkin, C.; Mackie, K. Distinct domains of the CB1 cannabinoid receptor mediate desensitization and internalization. J. Neurosci. 1999, 19, 3773–3780. [Google Scholar] [CrossRef] [Green Version]
  219. Delgado-Peraza, F.; Ahn, K.H.; Nogueras-Ortiz, C.; Mungrue, I.N.; Mackie, K.; Kendall, D.A.; Yudowski, G.A. Mechanisms of Biased β-Arrestin-Mediated Signaling Downstream from the Cannabinoid 1 Receptor. Mol. Pharmacol. 2016, 89, 618–629. [Google Scholar] [CrossRef] [Green Version]
  220. Roche, J.P.; Bounds, S.; Brown, S.; Mackie, K. A mutation in the second transmembrane region of the CB1 receptor selectively disrupts G protein signaling and prevents receptor internalization. Mol. Pharmacol. 1999, 56, 611–618. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram highlighting the role of nanoscale sub-compartmentalization of the dendritic spine in synaptic plasticity. The “basal” state can be depicted as a homeostatic equilibrium between synthesis, lateral diffusion, internalization, degradation, and recycling of neurotransmitter receptors at the dendritic spine. Activation of α7 nicotinic acetylcholine receptor (α7 nAChR) in the hippocampus by agonist on postsynaptic sites promotes LTP (left) by depolarizing the spine which induces glutamatergic GluA1 receptors to cluster at the PSD. This incorporation of GluA1 receptor molecules further contributes to calcium entry, thus strengthening synaptic transmission. The opposite phenomenon (LTD, right) is induced by activation of presynaptic cannabinoid receptors (CBRs): neurotransmitter release is inhibited, thereby weakening synaptic transmission, and GluA1 receptors are depopulated from the PSD. ABP: actin binding proteins; AN: anandamine.
Figure 1. Schematic diagram highlighting the role of nanoscale sub-compartmentalization of the dendritic spine in synaptic plasticity. The “basal” state can be depicted as a homeostatic equilibrium between synthesis, lateral diffusion, internalization, degradation, and recycling of neurotransmitter receptors at the dendritic spine. Activation of α7 nicotinic acetylcholine receptor (α7 nAChR) in the hippocampus by agonist on postsynaptic sites promotes LTP (left) by depolarizing the spine which induces glutamatergic GluA1 receptors to cluster at the PSD. This incorporation of GluA1 receptor molecules further contributes to calcium entry, thus strengthening synaptic transmission. The opposite phenomenon (LTD, right) is induced by activation of presynaptic cannabinoid receptors (CBRs): neurotransmitter release is inhibited, thereby weakening synaptic transmission, and GluA1 receptors are depopulated from the PSD. ABP: actin binding proteins; AN: anandamine.
Biomolecules 11 01697 g001
Figure 2. Schematic depiction of the enlargement and acquisition of mature mushroom-like shape of dendritic spines following induction of long-term potentiation (LTP) during synaptic plasticity.
Figure 2. Schematic depiction of the enlargement and acquisition of mature mushroom-like shape of dendritic spines following induction of long-term potentiation (LTP) during synaptic plasticity.
Biomolecules 11 01697 g002
Figure 3. Schematic diagram of a vomposite synapse summarizing the pre- and postsynaptic components participating in synaptic signaling dissected in this review. Glutamatergic (Glu) and cholinergic (α7 nAChR) activation promotes Ca2+ entry into the dendritic spine, inducing endocannabinoid (eCB) synthesis through hydrolysis of lipid precursors from the cell membrane. Cannabinoid receptor (CBR) activation in dendritic spines inhibits Rac1/WAVE/Arp 2/3 and limits the conversion of G actin to F-actin, whereas α7 nAChR promotes formation and maturation of dendritic spines through F-actin stability by activating the Rac1/WAVE/Arp2/3 signaling pathway. Phospholipase C (PLC) converts PIP2 into diacylglycerol (DAG) and in turn DAG lipase (DAGL) generates the eCB 2-AG. In parallel, phospholipase D (PLD) converts N-arachidonoyl phosphatidylethanolamine into the eCB AN. 2-AG and AN are liberated into the synaptic cleft and activate CBRs in the presynaptic compartment. Upon activation, CBRs stimulate Gi-protein and inhibit AC activity, membrane hyperpolarization ensues after the modulation of K+ and Ca2+ channels that inhibit neurotransmitter release from the presynaptic compartment. Finally, the mitogen-activated protein kinase (MAPK) pathway is stimulated. AC: adenyl cyclase; cAMP: cyclic AMP; ACh; acetylcholine; 2-AG, 2-arachidonyl glycerol.
Figure 3. Schematic diagram of a vomposite synapse summarizing the pre- and postsynaptic components participating in synaptic signaling dissected in this review. Glutamatergic (Glu) and cholinergic (α7 nAChR) activation promotes Ca2+ entry into the dendritic spine, inducing endocannabinoid (eCB) synthesis through hydrolysis of lipid precursors from the cell membrane. Cannabinoid receptor (CBR) activation in dendritic spines inhibits Rac1/WAVE/Arp 2/3 and limits the conversion of G actin to F-actin, whereas α7 nAChR promotes formation and maturation of dendritic spines through F-actin stability by activating the Rac1/WAVE/Arp2/3 signaling pathway. Phospholipase C (PLC) converts PIP2 into diacylglycerol (DAG) and in turn DAG lipase (DAGL) generates the eCB 2-AG. In parallel, phospholipase D (PLD) converts N-arachidonoyl phosphatidylethanolamine into the eCB AN. 2-AG and AN are liberated into the synaptic cleft and activate CBRs in the presynaptic compartment. Upon activation, CBRs stimulate Gi-protein and inhibit AC activity, membrane hyperpolarization ensues after the modulation of K+ and Ca2+ channels that inhibit neurotransmitter release from the presynaptic compartment. Finally, the mitogen-activated protein kinase (MAPK) pathway is stimulated. AC: adenyl cyclase; cAMP: cyclic AMP; ACh; acetylcholine; 2-AG, 2-arachidonyl glycerol.
Biomolecules 11 01697 g003
Figure 4. Diagrammatic depiction of dendritic spine remodeling during synaptic plasticity and associated neurotransmitter receptor clustering in nanodomains. (A) remodeling of the submembrane actin meshwork and incorporation of newly synthesized/laterally diffusing receptors from non-active areas of the spine head and neck into the PSD area, thereby increasing neurotransmitter receptor number at the site of contact with presynaptic boutons. At the crest of the spine receptors become entrapped by actin corrals, which might also present a lipid composition distinct from the bulk lipid bilayer. Golgi outpost in the dendrite and satellite Golgi outposts are found in the spine head, as well as smooth ER outposts. (B) Top view of the PSD. NMDARs (blue) are predominantly located at the center of the PSD in a single nanocluster, whereas AMPARs (red) are segregated into several nanodomains (sub-synaptic domains, SSD) surrounding the central NMDAR nanodomain. In contrast, mGluR5 (yellow) are aggregated into small clusters or homogeneously distributed at the PSD.
Figure 4. Diagrammatic depiction of dendritic spine remodeling during synaptic plasticity and associated neurotransmitter receptor clustering in nanodomains. (A) remodeling of the submembrane actin meshwork and incorporation of newly synthesized/laterally diffusing receptors from non-active areas of the spine head and neck into the PSD area, thereby increasing neurotransmitter receptor number at the site of contact with presynaptic boutons. At the crest of the spine receptors become entrapped by actin corrals, which might also present a lipid composition distinct from the bulk lipid bilayer. Golgi outpost in the dendrite and satellite Golgi outposts are found in the spine head, as well as smooth ER outposts. (B) Top view of the PSD. NMDARs (blue) are predominantly located at the center of the PSD in a single nanocluster, whereas AMPARs (red) are segregated into several nanodomains (sub-synaptic domains, SSD) surrounding the central NMDAR nanodomain. In contrast, mGluR5 (yellow) are aggregated into small clusters or homogeneously distributed at the PSD.
Biomolecules 11 01697 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vallés, A.S.; Barrantes, F.J. Nanoscale Sub-Compartmentalization of the Dendritic Spine Compartment. Biomolecules 2021, 11, 1697. https://doi.org/10.3390/biom11111697

AMA Style

Vallés AS, Barrantes FJ. Nanoscale Sub-Compartmentalization of the Dendritic Spine Compartment. Biomolecules. 2021; 11(11):1697. https://doi.org/10.3390/biom11111697

Chicago/Turabian Style

Vallés, Ana Sofía, and Francisco J. Barrantes. 2021. "Nanoscale Sub-Compartmentalization of the Dendritic Spine Compartment" Biomolecules 11, no. 11: 1697. https://doi.org/10.3390/biom11111697

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop