Next Article in Journal
Whole Genome Sequencing Solves an Atypical Form of Bardet–Biedl Syndrome: Identification of Novel Pathogenic Variants of BBS9
Previous Article in Journal
The Association of Genetic Markers Involved in Muscle Performance Responding to Lactate Levels during Physical Exercise Therapy by Nordic Walking in Patients with Long COVID Syndrome: A Nonrandomized Controlled Pilot Study
Previous Article in Special Issue
Brain-Specific Angiogenesis Inhibitor 3 Is Expressed in the Cochlea and Is Necessary for Hearing Function in Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neurotrophins and Their Receptors: BDNF’s Role in GABAergic Neurodevelopment and Disease

by
Carlos Hernández-del Caño
1,2,3,
Natalia Varela-Andrés
1,2,3,†,
Alejandro Cebrián-León
1,2,3,† and
Rubén Deogracias
1,2,3,*
1
Instituto de Neurociencias de Castilla y León (INCyL), 37007 Salamanca, Spain
2
Instituto de Investigación Biomédica de Salamanca (IBSAL), 37007 Salamanca, Spain
3
Departamento de Biología Celular y Patología, Facultad de Medicina, Universidad de Salamanca, 37007 Salamanca, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(15), 8312; https://doi.org/10.3390/ijms25158312
Submission received: 18 June 2024 / Revised: 21 July 2024 / Accepted: 25 July 2024 / Published: 30 July 2024

Abstract

:
Neurotrophins and their receptors are distinctly expressed during brain development and play crucial roles in the formation, survival, and function of neurons in the nervous system. Among these molecules, brain-derived neurotrophic factor (BDNF) has garnered significant attention due to its involvement in regulating GABAergic system development and function. In this review, we summarize and compare the expression patterns and roles of neurotrophins and their receptors in both the developing and adult brains of rodents, macaques, and humans. Then, we focus on the implications of BDNF in the development and function of GABAergic neurons from the cortex and the striatum, as both the presence of BDNF single nucleotide polymorphisms and disruptions in BDNF levels alter the excitatory/inhibitory balance in the brain. This imbalance has different implications in the pathogenesis of neurodevelopmental diseases like autism spectrum disorder (ASD), Rett syndrome (RTT), and schizophrenia (SCZ). Altogether, evidence shows that neurotrophins, especially BDNF, are essential for the development, maintenance, and function of the brain, and disruptions in their expression or signaling are common mechanisms in the pathophysiology of brain diseases.

1. Historical Perspective on Neurotrophins

Neurotrophins are a family of secreted proteins consisting of the nerve growth factor (NGF), the brain-derived neurotrophic factor (BDNF), neurotrophin-3 (NT-3), and neurotrophin-4/5 (NT-4/5).
Building upon the initial work of Spemann and Mangold regarding the “embryonic organizer” and embryonic development [1], the works of Viktor Hamburger and Rita Levi-Montalcini between the 1940s and the early 1950s led to the identification of NGF as the first factor promoting the survival and differentiation of sensory and sympathetic neurons [2,3,4].
In the early 1980s, the works by Yves-Alain Barde and Hans Thoenen led to the purification from pig brain of a factor that supported survival and growth of embryonic chick sensory neurons, later identified as BDNF [5]. The subsequent research revealed that BDNF plays a crucial role in promoting the survival and growth of neurons in the central nervous system and that BDNF disruption is involved in various neurological and psychiatric disorders [6]. These studies suggested that BDNF may play a crucial role in maintaining healthy brain functions and preventing the onset of mental illnesses, sparking a wave of research aimed at developing novel therapeutic interventions that target BDNF to treat these conditions, such as the development of therapies that aim to harness the power of BDNF to repair damaged neurons and restore brain functions [7].
In the early 1990s, a series of experiments that sought to isolate and characterize novel molecules analog to NGF and BDNF led to the discovery of NT-3 [8,9,10,11] and NT-4/5 [12,13,14,15].
In this review, we summarize the main characteristics and functions of the neurotrophins and their receptors. We also discuss the role of BDNF in the brain GABAergic inhibitory neurons, with special focus on two brain areas: the cerebral cortex and the striatum. Finally, we recapitulate the effects of BDNF signaling disruption in GABAergic neurons and its relationship with neurodevelopmental disorders.

2. Neurotrophins Expression, Synthesis, Structure, and Secretion in Brain

2.1. Expression in the Developing Brain

During brain development, neurotrophins exhibit distinct spatial and temporal patterns of RNA expression and protein distribution (Figure 1). In the embryonic rat brain, the most abundant neurotrophin transcripts are Nt-3 and Nt-4/5, while Ngf expression is intermediate, and Bdnf is scarcely detected. These patterns shift after birth. In postnatal development, while Ngf RNA levels are relatively constant in the entire brain, Nt-3 expression decreases while Bdnf expression increases until adulthood. Nt-4/5 RNA levels decay to intermediate levels and then remain stable (Figure 1A) [10,15].
In rodents, it has been observed that the expression levels of different neurotrophins vary not only through developmental stages but also between brain regions (Figure 1B). For instance, in newborn rats, Nt-3 expression is higher in the cerebellum, hippocampus, and cortex, which are more immature areas at this stage, whereas in the adult it remains high only in the hippocampus [10]. Conversely, Bdnf expression in the newborn is primarily located in areas that mature earlier, such as the hindbrain, midbrain, and diencephalon, as well as in the hippocampus. In the adult brain, the highest levels of Bdnf RNA are found in the hippocampus, with abundant levels in the neocortex, diencephalon, midbrain, and cerebellum and decreasing levels in the hindbrain and spinal cord [10,15,16,17]. It is noteworthy that neither Nt-3 nor Bdnf are expressed in the striatum during both the newborn and adult stages [10]. Finally, Nt-4/5 displays a similar expression pattern in different brain regions along postnatal development, except in the cerebellum, where it shows several expression peaks [15].
Figure 1. Schematic representations of neurotrophin RNA and protein relative level distribution change along brain development and comparing between species. (A): Neurotrophin RNA levels in whole brain lysates along rat development. Data acquired from northern blot assays from reference [10] and RNAse protection assays from reference [15] using probes against the coding sequence (CDS) of each neurotrophin. E: embryonic gestation day; P: postnatal day. (B): Ngf, Bdnf, and Nt-3 RNA levels in different brain regions of newborn (NB) and adult (AD) rats, and Nt-4/5 along postnatal rat development in different brain regions. Data acquired from northern blot assays from reference [10] and RNase protection assays from reference [15], using probes against the CDS of each neurotrophin. (C): NGF, BDNF, NT-3, and NT-4/5 protein distribution in different brain areas from adult rodents and monkeys. Rodent data were acquired from both immunohistochemistry and protein quantification experiments from references [18,19,20,21]. Monkey data were obtained from immunohistochemistry experiments from reference [22]. In all cases, used antibodies do not discriminate between precursors and mature neurotrophins. Darker colors mark higher RNA or protein levels, relativized to the maximum expressing point. N/D: data not available.
Figure 1. Schematic representations of neurotrophin RNA and protein relative level distribution change along brain development and comparing between species. (A): Neurotrophin RNA levels in whole brain lysates along rat development. Data acquired from northern blot assays from reference [10] and RNAse protection assays from reference [15] using probes against the coding sequence (CDS) of each neurotrophin. E: embryonic gestation day; P: postnatal day. (B): Ngf, Bdnf, and Nt-3 RNA levels in different brain regions of newborn (NB) and adult (AD) rats, and Nt-4/5 along postnatal rat development in different brain regions. Data acquired from northern blot assays from reference [10] and RNase protection assays from reference [15], using probes against the CDS of each neurotrophin. (C): NGF, BDNF, NT-3, and NT-4/5 protein distribution in different brain areas from adult rodents and monkeys. Rodent data were acquired from both immunohistochemistry and protein quantification experiments from references [18,19,20,21]. Monkey data were obtained from immunohistochemistry experiments from reference [22]. In all cases, used antibodies do not discriminate between precursors and mature neurotrophins. Darker colors mark higher RNA or protein levels, relativized to the maximum expressing point. N/D: data not available.
Ijms 25 08312 g001

2.2. Synthesis, Secretion, and Protein Levels

Neurotrophins, like most of growth factors, are initially synthesized in the rough endoplasmic reticulum as precursors called proneurotrophins, composed of a N-terminal prodomain and a C-terminal mature domain. These proneurotrophins are then packaged into secretory vesicles for processing by convertase family proteases, leading to the production of mature neurotrophins. In the case of BDNF, it is initially synthesized and glycosylated as a precursor (pro-BDNF), which then moves into the Golgi apparatus. Here, it may undergo endoproteolytic cleavage, promoting its targeting to secretory granules, or it may be directly secreted as a proneurotrophin, which can be processed by extracellular proteases or act as an independent signal [23].
As secreted molecules, protein levels and distribution do not directly correspond to their expression pattern. Also, both the locations and the protein levels of neurotrophins differ between species (Figure 1C) [17,18,19,20,21,22]. Adult rodents show the highest immunoreactivity for NGF in the cortex, the hippocampus, and the cerebellum [19], whereas in the rhesus monkey the highest immunoreactivity for NGF is detected the striatum, the midbrain, and the hindbrain [22]. BDNF immunoreactivity is higher in the rodent neocortex, hippocampus, and hindbrain [20,21], while in monkeys the highest immunoreactivity is located in the hippocampus, hindbrain, and thalamus/hypothalamus [22]. NT-3 levels are higher in the rat hippocampus, thalamus, and cerebellum [18,21], while in the macaque they are higher in the hippocampus, midbrain, and hindbrain [22]. NT-4/5 levels are especially abundant in the cerebellum in rodents [21], whereas in the monkey they are low throughout the whole brain except in the midbrain. Notably, both NT-3 and NT-4/5 are absent in the macaque thalamus [22].

2.3. BDNF Expression and Protein Location during Development

During development, the expression of BDNF is dynamically regulated in a region-specific manner. In both mice and human the gene encoding BDNF contains several 5′-non-coding exons driven by distinct promoter regions, resulting in a common protein-coding 3′ exon-mRNA [17,24,25]. These mRNA variants create a “spatial code” with distinct subcellular distribution of these transcripts in neuronal and non-neuronal cells, implicating them in many different functions [17,26]. For example, regulation of dendrite complexity [27], thermogenesis, body weight [28], aggression in male mice [29], impaired maternal care [30], and impaired inhibitory synapse formation [31,32] are such functions.
In mice, Bdnf expression increases throughout development until adulthood, with the mouse hippocampus particularly enriched in Bdnf transcripts during postnatal development. In humans, the highest expression peak is observed in the thalamus during infancy, with a similar progression in the amygdala, cerebellum, and dorsolateral prefrontal cortex during development. However, almost no mRNA can be detected in the human striatum at any developmental stage. The levels, proportions, and distributions of the various BDNF mRNA variants change differently during brain development across species (Figure 2A) and even between mouse strains [17].
BDNF protein levels also increase during postnatal development, aligning with the maturation times of different brain areas (Figure 2B). Early maturing regions such as the thalamus, hypothalamus, midbrain, and hindbrain exhibit higher BDNF levels at earlier time points during postnatal development, while other regions like the olfactory bulb, neocortex, hippocampus, and striatum begin to show increased BDNF levels from approximately P10. The cerebellum displays a later increment of BDNF levels, being one of the last brain areas to develop [17,33].
The subcellular distribution of BDNF varies among brain regions (Figure 2C). In adult rats, BDNF immunoreactivity is higher inside the cell bodies in neurons located in the neocortex and cerebellum, while in the hippocampus, striatum, thalamus, hypothalamus, midbrain, and hindbrain, it is more prominent in the fibers [20]. This distribution, akin to the subcellular distribution of the alternative mRNAs [26], also represents a “spatial code” that likely regulates BDNF functions. Whether this “code” remains constant during development or changes with age to exert different roles is still unknown.
Figure 2. Schematic representations of BDNF mRNA and protein distribution in the developing brain. (A): BDNF CDS-mRNA distribution in different brain areas in mouse (left) and human (right) along development, according to RNAseq data from reference [17]. Darker colors mark higher mRNA levels, relativized to the maximum expressing point. (B): BDNF protein levels along mouse postnatal development in different brain areas, considering the mean of the detected protein levels by western blot in both BALB/C and C57BL6/J mice in reference [17]. (C): BDNF subcellular distribution in different brain areas in the adult rat brain, according to reference [20]. E: mouse embryonic stage; P: mouse postnatal day; PCW: human postcoital week. N/D: data not available.
Figure 2. Schematic representations of BDNF mRNA and protein distribution in the developing brain. (A): BDNF CDS-mRNA distribution in different brain areas in mouse (left) and human (right) along development, according to RNAseq data from reference [17]. Darker colors mark higher mRNA levels, relativized to the maximum expressing point. (B): BDNF protein levels along mouse postnatal development in different brain areas, considering the mean of the detected protein levels by western blot in both BALB/C and C57BL6/J mice in reference [17]. (C): BDNF subcellular distribution in different brain areas in the adult rat brain, according to reference [20]. E: mouse embryonic stage; P: mouse postnatal day; PCW: human postcoital week. N/D: data not available.
Ijms 25 08312 g002

3. Neurotrophin Receptors

Neurotrophins exert their effects by binding with high affinity to specific cell surface receptors known as tropomyosin receptor kinases (Trks), namely TrkA, TrkB, and TrkC. These tyrosine kinase receptors dimerize upon ligand binding and undergo transphosphorylation, where the intracellular kinase domain of the receptor phosphorylates itself and promotes the activation of intracellular protein effectors [34,35]. Alternatively, the non-processed neurotrophins or proneurotrophins, as well as the mature fully processed neurotrophins, bind to the low-affinity p75 neurotrophin receptor (p75NTR) [36,37,38,39,40].

3.1. Trk Receptors

Trk receptors are expressed differentially during development (Figure 3B) [41] and in different subsets of neurons, playing a wide range of roles after neurotrophin binding. Surprisingly, Trk receptors can also be activated in the absence of neurotrophins by trans-activation mechanisms [42,43,44,45,46,47,48]. Therefore, it is not surprising that some phenotypes observed in Trk-deficient mice [49] do not totally phenocopy the ones observed in neurotrophin-deficient mice [50,51,52,53].
The structure of various Trk receptors is very similar (Figure 3A): an extracellular domain composed of one cysteine-rich domain followed by three leucine-rich motifs and another cysteine-rich domain, two immunoglobulin-like C2 domains, a transmembrane domain (which is responsible for dimerization after neurotrophin binding), and a cytoplasmic tyrosine kinase domain that guides the downstream signaling. However, the existence of several splicing isoforms of all Trk receptors that differ in the cytoplasmic domain has been reported [34].
Upon neurotrophin binding, Trk receptors activate three different signaling pathways, the PLCγ/IP3, the PI3K/Akt, and the MAPK pathways, which mediate multiple and specific neurotrophin-induced biological responses [54].
In general, these responses are modulation of gene expression, presynaptic neurotransmitter release, postsynaptic neurotransmitter receptor exposure and function, ion channel activity and conductance, regulation of dendritic growth and morphology, and spine maturation [55].
In addition to their canonical direct activation, Trk receptors can be transactivated through alternative neurotrophin-independent pathways [44]. The transactivation of Trk receptors was initially observed in studies investigating the signaling pathways of the adenosine and pituitary adenylate cyclase-activating polypeptide (PACAP) through G-protein-coupled receptors [42,43]. These studies provided early insights into the ability of Trk receptors to be activated through non-canonical pathways. More recently, in vivo studies have demonstrated the transactivation of TrkB and TrkC by the epidermal growth factor (EGF) to drive the migration of newborn mouse cortical neurons [47]. Additionally, transactivation of TrkB has been observed in response to glucocorticoids [45], zinc [46], and, more recently, to oxytocin [48].
Among the different Trk isoforms (Figure 3A), the most well studied is TrkB-T1, a truncated TrkB isoform lacking the intracellular tyrosine kinase domain but with a small aminoacidic tail that transduces BDNF signaling into the cytosol [34,56]. TrkB-T1 presents a unique cytoplasmic tail of 11 amino acids fully conserved across species, suggesting a potential important evolutionary biological role [57]. TrkB-T1 is mainly, but not exclusively, expressed in astrocytes [58], and the function of its cytoplasmic tail is still not fully understood. However, recent research has implicated TrkB-T1 in various biological processes, including neural development, early embryonic central nervous system and mesenchymal development, glioma biology, neuropathic pain, regulation of cell morphology, and the functionality of glycine transporters [58,59]. In addition, TrkB-T1 has been shown to exhibit an important neuroprotectant function [60,61], and it has been reported to interact with p75NTR in modulating functional and structural plasticity in hippocampal neurons [62].

3.2. The p75 Neurotrophin Receptor

The receptor p75NTR plays a significant role in binding NGF, BDNF [36], NT-3 [37], and NT-4/5 [38]. Structurally, it consists of an extracellular domain with four cysteine-rich subdomains that binds the four mature neurotrophins with similar affinity. Its intracellular portion contains a chopper domain and death domain (Figure 4) [40].
During the development of the central and peripheral nervous system, p75NTR expression is high, particularly during synaptogenesis and developmental cell death. However, its expression is downregulated in adulthood (Figure 5C) [17]. Nonetheless, following injury or disease, p75NTR expression can increase once again [63].
It is worth noting that proneurotrophins can be directly secreted in disease states and interact with p75NTR [23]. Increased levels of pro-NGF have been observed in the frontal and parietal cortex of Alzheimer’s patients compared to controls [64,65]. Conversely, in a cleavage-resistant pro-BDNF knockin mouse, secretion of pro-BDNF has been shown to negatively regulate synapse plasticity and transmission [66].
The interaction between proneurotrophins and p75NTR can induce apoptosis, even at sub-nanomolar concentrations. This occurs through the formation of a heterodimer between p75NTR and the type I transmembrane protein sortilin. Sortilin recognizes the proneurotrophin prodomain, while p75NTR recognizes the domain corresponding to the mature neurotrophin. This apoptotic mechanism involves the activation of the JNK-p53 apoptotic pathway and subsequent procaspase-mediated cleavage of different substrates [67,68,69,70]. Furthermore, p75NTR has been associated with cytoskeletal reorganization through interactions with proteins such as RhoA [71] and fascin [72]. p75NTR can also be cleaved by the α- and γ-secretases, splitting it into its extracellular (p75ECD) and intracellular (p75ICD) domains. p75NTR then interacts through its chopper domain with the intracellular domain of TrkA receptors establishing a tetramer (Figure 4). This interaction allosterically modulates TrkA binding affinity for NGF, increasing it and mediating effects on cell survival and differentiation [40]. It also plays a role in neuronal morphology, including dendritic arborization, axonal pruning, and neurite retraction [73,74,75]. Neurotrophins, including both proneurotrophins and mature forms, have specific effects on synaptic plasticity. The secretion of pro-BDNF and mature BDNF is influenced by different neuronal activities, with pro-BDNF being predominant in low-frequency stimulation associated with long-term depression (LTD), and mature BDNF being released following high-frequency stimulation associated with long-term potentiation (LTP) [76,77]. In the CA1 area of the hippocampus, pro-BDNF enhances LTD, while mature BDNF is required for the maintenance of LTP [55,78]. Moreover, neurotrophins and p75NTR are involved in the control of cell cycle and proliferation. Neurons are typically in a quiescent state, but cell cycle reactivation can lead to apoptosis or transition to tetraploid neurons. Activation of cell proliferation factors or neurotrophic factor deprivation can trigger this process, and in disease states like Alzheimer’s, proNGF activation ultimately results in neuronal death [79,80].

4. Role of BDNF in GABAergic Neurons

BDNF plays distinct roles in glutamatergic and GABAergic neurons. The BDNF/TrkB signaling pathway in the hippocampus is critical for proper synaptic function, and disruptions in this pathway have been implicated in various diseases [81]. The dysregulation of BDNF signaling can have significant consequences for synaptic plasticity and overall hippocampal function. In the hippocampus, BDNF is involved in mediating the formation of functional synapses in both types of neurons [82]. However, in hippocampal GABAergic populations, BDNF not only influences synapse formation but also modulates their growth, differentiation, and synaptic plasticity [53,82,83,84]. Interestingly, in hippocampal glutamatergic neurons, BDNF does not control neuronal growth but rather the neurotrophin NT-3 takes on that role [82]. This highlights the multifaceted role of BDNF in shaping the development and function of GABAergic neurons in the hippocampus and demonstrates the specificity of neurotrophin regulation in different neuronal populations within the same brain region.
Furthermore, BDNF also plays a role in specifically regulating GABAergic populations in other brain areas such as the cortex or the striatum.

4.1. Role of BDNF in the GABAergic Cortex

Neurogenesis in rodents begins around embryonic stage E9.5 with the division of neuroepithelial cells, leading to the formation of radial glial cells and basal progenitors [85]. These precursor cells then undergo radial migration to form the cortical layers, with deeper layers migrating first and outer layers arriving last.
During this process, BDNF and NT-3 expression in the brain is minimal [10]. However, the signaling of TrkB and TrkC receptors, transactivated by the epidermal growth factor receptor (EGFR), regulates this migration [47]. The addition of exogenous BDNF during these embryonic stages leads to the induction of neuronal bone morphogenetic protein 7 (BMP7) expression through the BDNF/MAPK/ERK1/2 pathway, resulting in premature and incorrect differentiation of radial glial cells and impaired neuronal migration [86].
The integration of interneurons in the cortex occurs primarily during early postnatal development and involves programmed cell death and circuit refinement [87,88]. BDNF levels begin to rise at this point [10], and BDNF is known to play a crucial role in the maturation of GABAergic interneurons. Specifically, BDNF is required for inhibitory cortical neurons’ dendritic development but not for that of excitatory neurons [89]. This effect on inhibitory neurons is mainly seen during the late period of corticogenesis, where BDNF supports the maintenance of neuron size and dendrite structure [90].
Among the interneurons, parvalbumin-expressing neurons appear to be particularly dependent on BDNF and TrkB signaling, which regulates parvalbumin expression during early postnatal development in the visual and prefrontal cortex [91,92]. GABA transmission in cortical interneurons is modulated by BDNF, which influences the transcription of the GABA-synthesizing enzyme GAD65 in a MAPK/ERK/CREB-dependent manner [93]. On the other hand, it has been observed that hyperactivation of the MEK1/ERK1/2 signaling pathway leads to specific defects in cortical parvalbumin neurons, resulting in impaired behavioral inhibition [94]. TrkB signaling also regulates the dynamics of parvalbumin cortical neurons, modulating their excitability and the number and strength of synapses they make with pyramidal neurons [95]. TrkB, activated by both BDNF and NT-4, also mediates the expression of Kv3.1b and Kv3.2, two essential potassium channels for the activity of fast-spiking basket and chandelier cells [96].
The incoming synapses onto parvalbumin neurons are heavily regulated by the perineuronal net (PNN), which influences their function [97]. The development of the PNN depends on BDNF and JNK signaling, and disruptions in the PNN have been linked to schizophrenia [98].
p75NTR signaling also plays a significant role in the development and function of parvalbumin neurons, as its expression is observed in early cortical progenitors and persists in post-mitotic cortical neurons and migrating neurons within the cortical plate [99]. Conditional deletion of p75NTR in mice results in impairments in the development of cortical interneurons and upper-layer pyramidal neurons, leading to the loss of cortical layer thickness and a decrease in the number of parvalbumin, somatostatin, neuropeptide Y, and calretinin-positive cells [99]. p75NTR is also required for the survival of cortical neuron progenitors and the production of later-born neurons. In parvalbumin neurons, p75NTR regulates the timing of maturation and connectivity in the visual cortex, affecting the development of their perineuronal nets and dynamically regulating them through pro-BDNF-dependent plasticity mechanisms [100]. Deletion of p75NTR causes perineuronal net aggregation in the prefrontal cortex, which can be rescued by reintroducing p75NTR in preadolescent, but not postadolescent, prefrontal parvalbumin neurons [101]. These results show that p75NTR signaling not only is relevant in disease states but is also essential for proper cortical wiring, indicating as well that more research needs to be conducted regarding p75NTR physiological function in the normal development of the cortex and other brain areas.

4.2. Role of BDNF in the Striatum

The striatum, a prominent GABAergic region within the basal ganglia, is primarily composed of Medium Spiny Neurons (MSNs) derived from the lateral ganglionic eminence, along with several interneuron populations from various sources [102].
The MSNs receive glutamatergic inputs from cortical and thalamic areas, as well as dopaminergic inputs from the substantia nigra pars compacta and the ventral tegmental area [103]. The MSNs also undergo local modulation by interneurons and other MSNs [104].
There are functionally distinct populations of MSNs, expressing either the dopamine receptor D1 or D2 [105]. The MSNs expressing the D1 receptor project onto the substantia nigra, while the D2 project into the globus pallidus, forming the direct and indirect dopaminergic pathways, respectively [104,106].
Functionally, the striatum is mainly involved in the cognitive planning of purposive motor acts [107], but it also participates in other functions like decision-making, reward processing, and response inhibition, among others [108].
In the adult striatum, while Bdnf mRNA is absent [10], BDNF protein levels are high as it is anterogradely transported to the striatum from the cerebral cortex, substantia nigra, amygdala, and thalamus [17,20,109]. BDNF secretion in the corticostriatal circuitry exhibits a specific distribution pattern. Through the utilization of retrograde viral tracing, it has been observed that neurons expressing Bdnf originating from the medial prefrontal cortex primarily project onto the dorsomedial striatum. Conversely, neurons originating from the motor and the agranular insular cortices extend their axons to the dorsolateral striatum. Moreover, the outputs from the orbitofrontal cortical neurons are directed toward the dorsal striatum, with the specific target location being dependent on the mediolateral and rostrocaudal positioning of these neurons. Notably, distinct populations of cortical neurons have been identified to release BDNF at their axon terminals, thereby influencing BDNF/TrkB signaling within different areas of the striatum [110]. This corticostriatal parcellation, in conjunction with striatal outputs, also plays a crucial role in establishing the functional subdivisions of the striatum into limbic, associative, and sensorimotor areas [111].
Furthermore, the presence of TrkB in the membrane of the MSNs is intrinsically modulated. Activation of the dopamine indirect pathway causes the retraction of TrkB from the plasma membrane [112], whereas the direct pathway enhances TrkB’s sensitivity to BDNF by modulating TrkB presence on the cell surface [113].
Conditional BDNF knockout mice have revealed the importance of BDNF in proper striatal outgrowth, MSN arborization, and functional development. Additionally, BDNF/TrkB signaling controls the number of newborn striatal neurons and supports the survival of immature MSNs [53,114,115,116].
BDNF downstream signaling also regulates striatal synaptic dynamics, influencing synapse-related genes, drug response, and addiction processes [117]. In vivo, BDNF-mediated dopamine potentiation in the striatum requires specific pathway activations, such as PI3K and MAPK, but not PLCγ [118]. However, PLCγ1 plays a crucial role in establishing the dopaminergic system and regulates dopamine release input into the striatum [119,120]. PLCγ hydrolyses PIP2 into DAG and IP3, both of which heavily modulate synaptic transmission in striatal neurons, affecting MSNs’ output neuron activity and voltage-gated sodium channels within MSNs [121,122].
Regarding TrkB truncated isoforms, TrkB-T1 is not expressed in the striatum [123] but modulates corticostriatal transmission, inhibiting BDNF signaling by NT-4/5 [124]. In vitro, excitotoxic stimulation downregulates TrkB-FL via calpains and upregulates truncated TrkB protein levels through activation of its transcription and translation [125]. ProBDNF/p75NTR signaling is also essential in the striatum, mediating the survival of neuronal progenitors and the development of the basal ganglia. Disruption of p75NTR has been linked with Huntington’s disease, as it affects corticostriatal synaptic LTP [126] and the survival of striatal neurons [127]. p75NTR alterations are also related with Parkinson’s disease, alcohol abuse, cocaine sensitization, ischemic trauma, and HIV-associated neurological disorder [99,128,129]. However, the underlying downstream molecular mechanisms guiding these roles remain to be fully elucidated.

5. BDNF on Neurodevelopmental Disorders with Impairments in GABAergic Neurons

Dysfunctions in the formation of the GABAergic system are linked to the appearance of neurodevelopmental disorders, mainly due to a loss in the balance between excitation and inhibition in the brain [130]. Autism spectrum disorder (ASD), Rett syndrome (RTT), and schizophrenia (SCZ) are three neurodevelopmental disorders that show this imbalance derived from impairments within the GABAergic brain network and have been related with BDNF-signaling disruptions.

5.1. BDNF in Autism Spectrum Disorder

Autism spectrum disorder (ASD) is characterized by impairments in social interaction, communication, and repetitive behaviors, often accompanied by co-occurring symptoms such as dyskinesia, speech delay, sleep disorder, anxiety, and epilepsy in children and depressive symptoms in adolescents and adults [131,132]. The etiology of ASD is complex, involving abnormal interactions among genetic and environmental factors, leading to epigenetic changes that modify gene transcription and cell function [131].
Neurodevelopmental alterations associated with ASD involve changes in molecular processes, including chromatin remodeling and Ca2+ and Wnt signaling [131]. Wnt signaling, critical for cell fate determination and neural development, plays a key role in normal brain growth, neuron proliferation, and connectivity [133,134]. These alterations modify BDNF expression and signaling in the brain, as BDNF expression is regulated by neural activity and Ca2+ signaling via transcription factors such as CREB, NFAT, and MEF2, which are dysregulated in ASD [131,135,136].
In animal models with autism-like behavior, such as BTBR mice and neuroligin 3 (NLG3) loss-of-function mutants, impairments in GABAergic hippocampal synaptic function have been found and directly related to BDNF/TrkB signaling impairments [137,138].
Epigenetic mechanisms, including DNA methylation and histone modification, are also involved in ASD, affecting social behavior regulation and GABAergic transmission [131,139]. Enhanced binding of MeCP2 to Gad1 and Gad2 promoter regions, reducing the expression of GABA synthesizing enzymes, has been observed in ASD patients, impairing GABAergic transmission [140]. Altered expression of the oxytocin receptor, which can transactivate TrkB [48], has also been linked to ASD, showing high levels of methylation of its promoter [131].
Additionally, the BDNF Val66Met mutation, which diminishes activity-dependent BDNF signaling [141,142], has been associated with autism-like social deficits in mice, with differences between males and females [143].
The relationship between BDNF alterations and ASD is not entirely clear, as it may involve both causative and consequential factors. Neurodevelopmental alterations prior to BDNF’s functional roles in neuron maturation may lead to BDNF misregulation as a consequence of the disorder, potentially driving the appearance of various co-occurring symptoms. Conversely, disruptions in BDNF signaling can contribute to some of the behavioral deficits in ASD, serving as one of the causative factors of the disease, particularly through its role in GABAergic neuron maturation, given the strong association between GABAergic dysfunction and the appearance of ASD traits [131,140]. The intricate interplay between BDNF signaling and ASD underscores the multifaceted nature of the disorder, involving various genetic, epigenetic, and molecular mechanisms, which, to this day, are not fully understood.

5.2. BDNF in Rett Syndrome

Rett syndrome (RTT) is a severe neurodevelopmental disorder that primarily affects females, with rare cases affecting males. In males, RTT leads to rapid fatal neonatal encephalopathy, while in females, it involves developmental regression, including the loss of acquired language and social skills, motor impairments, autonomic dysfunction, and seizures [144,145,146].
Classical RTT is caused by loss-of-function mutations in the Methyl-CpG Binding protein 2 (MeCP2), while atypical RTT can result from mutations in other genes such as CDKL5 and FOXG1 [147].
MeCP2 is an epigenetic regulator that plays a crucial role in reading DNA methylation to mediate DNA binding of the repressors NCOR1 [148,149] and Sin3A [150,151] or of the activator transcription factor CREB [150].
Through these interactions, MeCP2 modulates BDNF expression [145], with BDNF also regulating MeCP2 phosphorylation. BDNF signaling in cortical neurons mediates several MeCP2 phosphorylations [152], influencing cortical dendritic branching, activity, and sustained effects of antidepressants [148,153,154,155].
Dysregulation of the brain’s excitatory/inhibitory balance, mainly caused by impairments in the brain’s GABAergic network, is a physiological hallmark of RTT etiology [156]. MeCP2 deletion specifically from GABAergic neurons replicates most of the phenotypical features of MeCP2 full KO mice, while restoration of MeCP2 only in GABAergic neurons rescues ataxia, apraxia, and social abnormalities, and extends lifespan [146].
Given the importance of GABAergic network impairments in RTT etiology and the molecular relationship between BDNF and MeCP2, BDNF has become a target for RTT therapies. BDNF overexpression in postnatal excitatory forebrain and hippocampal neurons has shown promise in reverting some RTT characteristics, highlighting the neuroprotective function of BDNF [157,158].
Pharmacological approaches that enhance BDNF-dependent signaling have shown promise as therapy for RTT. TrkB agonists like LM22A-4 [159] and 7,8-DHF [160] have been able to revert some RTT symptoms in animal models, including restoration of hippocampal synaptic plasticity and object location memory [161]. Additionally, Fingolimod, a compound that stimulates the MAPK/ERK signaling pathway through modulation of Sphingosine-1 phosphate receptors, has shown promise in improving RTT-like features in MeCP2 KO mice [162], although it failed to provide clear results in clinical trials [163]. Recently, the FDA approved the first Rett syndrome treatment, Trofinetide, an endogenous tripeptide of the N-terminal domain of IGF-1 [164].

5.3. BDNF in Schizophrenia

Schizophrenia, a neurodevelopmental psychiatric disorder that affects almost 1% of the general population [165] manifests with positive and negative symptoms, including psychosis, decreased expression of emotions, catatonia, and social and occupational decline [111,166]. The “mesolimbic hypothesis” is a well-established explanation for schizophrenia pathophysiology, attributing impairments in dopaminergic modulation in mesolimbic areas, particularly the striatum, to the disorder’s etiology. Altered dopamine projections onto the striatum from various brain regions contribute to its role as an integrative signaling hub [111,165].
The prefrontal cortex is also heavily implicated in schizophrenia onset, where impairments in parvalbumin neurons have been linked to schizophrenia-like phenotypes in mice [111,167,168]. Despite BDNF playing a critical role in the development and functionality of cortical parvalbumin neurons, the relationship between BDNF signaling impairments in these neurons and schizophrenia remains largely uncharacterized. Research has demonstrated that dopamine D1 receptor transmission in the dorsolateral prefrontal cortex is elevated in schizophrenia patients, regulating BDNF expression [169,170]. Furthermore, BDNF is secreted from these neurons to the striatum, where it regulates the expression of the dopamine receptor D3, which is overexpressed in schizophrenia patients [110,171,172].
BDNF haploinsufficiency causes schizophrenia-like phenotypes, which can be rescued under rich-environment conditions that promote BDNF expression and secretion in the brain [173], establishing a clear relationship between BDNF function impairments and the development of schizophrenia symptoms, influenced by environmental conditions.
Consequently, BDNF has been extensively studied in both a therapeutic target and a biomarker for schizophrenia [165,174]. Polymorphisms in BDNF and altered post-mortem BDNF brain levels have been identified in schizophrenia patients [174], and animal models have shown both decreased and increased BDNF levels in different brain regions [175,176]. However, using serum or plasma levels of BDNF as a biomarker presents challenges due to the release of BDNF from human blood platelets [177,178], leading to varied results in studies analyzing BDNF levels in schizophrenia patients [165,179].

6. Concluding Remarks

Neurotrophins and their receptors play a critical role in neural development and function. They exhibit distinct temporal and spatial expression patterns that not only are different at the species level, but also during their developmental stages and in brain areas.
Among the neurotrophin family, BDNF has emerged as a key element preserving neuronal health. The fine tuning of BDNF expression and its protein distribution constitutes a “spatial code” that stablishes the background for its actions through TrkB interaction. BDNF/TrkB signaling is key for GABAergic neuron development, maintenance, and function, being crucial for maintaining the excitatory/inhibitory equilibrium in the brain.
Imbalances in both BDNF expression and its signaling cause significant implications of this neurotrophin in various diseases. Different studies support the notion that impairments in the GABAergic system driven directly or indirectly by BDNF may play an important role in the pathophysiology of neurodevelopmental disorders like autism spectrum disorder, Rett syndrome, and schizophrenia, as mentioned.
Despite the fact that distinct effects of BDNF signaling in the brain are well characterized, the underlying mechanisms intrinsic for each neuronal population that modulate the effects of this signaling remain quite unknown. Further research is needed to reveal the complex but precise molecular interactions that drive BDNF expression and activity, possibly leading to the discovery of novel therapeutic targets that overcome the challenge of neurodevelopmental diseases.

Author Contributions

Writing—original draft preparation, C.H.-d.C. and R.D.; writing—review and editing, C.H.-d.C., N.V.-A., A.C.-L. and R.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grants from the Spanish Ministry of Science and Innovation to R.D. (RYC2018/205215-I, PID2020-113086RB-I00 and CNS2022-136048). N.V.-A. has a contract of the Program III: Grants for predoctoral contracts 2021 co-funded by the Universidad de Salamanca and Banco Santander. A.C.-L. and C.H.-d.C.’s contracts are associated to PID2020-113086RB-I00 and CNS2022-136048 grants, respectively.

Data Availability Statement

Data were obtained from and are available in the references indicated in each figure.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Spemann, H.; Mangold, H. Induction of Embryonic Primordia by Implantation of Organizers from a Different Species. 1923. Int. J. Dev. Biol. 2001, 45, 13–38. [Google Scholar] [PubMed]
  2. Hamburger, V.; Levi-Montalcini, R. Proliferation, Differentiation and Degeneration in the Spinal Ganglia of the Chick Embryo under Normal and Experimental Conditions. J. Exp. Zool. 1949, 111, 457–501. [Google Scholar] [CrossRef] [PubMed]
  3. Levi-Montalcini, R.; Hamburger, V. Selective Growth Stimulating Effects of Mouse Sarcoma on the Sensory and Sympathetic Nervous System of the Chick Embryo. J. Exp. Zool. 1951, 116, 321–361. [Google Scholar] [CrossRef] [PubMed]
  4. Levi-Montalcini, R.; Hamburger, V. A Diffusible Agent of Mouse Sarcoma, Producing Hyperplasia of Sympathetic Ganglia and Hyperneurotization of Viscera in the Chick Embryo. J. Exp. Zool. 1953, 123, 233–287. [Google Scholar] [CrossRef]
  5. Barde, Y.A.; Edgar, D.; Thoenen, H. Purification of a New Neurotrophic Factor from Mammalian Brain. EMBO J. 1982, 1, 549–553. [Google Scholar] [CrossRef]
  6. Lima Giacobbo, B.; Doorduin, J.; Klein, H.C.; Dierckx, R.A.J.O.; Bromberg, E.; De Vries, E.F.J. Brain-Derived Neurotrophic Factor in Brain Disorders: Focus on Neuroinflammation. Mol. Neurobiol. 2019, 56, 3295–3312. [Google Scholar] [CrossRef] [PubMed]
  7. Miranda-Lourenço, C.; Ribeiro-Rodrigues, L.; Fonseca-Gomes, J.; Tanqueiro, S.R.; Belo, R.F.; Ferreira, C.B.; Rei, N.; Ferreira-Manso, M.; De Almeida-Borlido, C.; Costa-Coelho, T.; et al. Challenges of BDNF-Based Therapies: From Common to Rare Diseases. Pharmacol. Res. 2020, 162, 105281. [Google Scholar] [CrossRef]
  8. Hohn, A.; Leibrock, J.; Bailey, K.; Barde, Y.-A. Identification and Characterization of a Novel Member of the Nerve Growth Factor/Brain-Derived Neurotrophic Factor Family. Nature 1990, 344, 339–341. [Google Scholar] [CrossRef]
  9. Jones, K.R.; Reichardt, L.F. Molecular Cloning of a Human Gene That Is a Member of the Nerve Growth Factor Family. Proc. Natl. Acad. Sci. USA 1990, 87, 8060–8064. [Google Scholar] [CrossRef]
  10. Maisonpierre, P.C.; Belluscio, L.; Friedman, B.; Alderson, R.F.; Wiegand, S.J.; Furth, M.E.; Lindsay, R.M.; Yancopoulos, G.D. NT-3, BDNF, and NGF in the Developing Rat Nervous System: Parallel as Well as Reciprocal Patterns of Expression. Neuron 1990, 5, 501–509. [Google Scholar] [CrossRef]
  11. Rosenthal, A.; Goeddel, D.V.; Nguyen, T.; Lewis, M.; Shih, A.; Laramee, G.R.; Nikolics, K.; Winslow, J.W. Primary Structure and Biological Activity of a Novel Human Neurotrophic Factor. Neuron 1990, 4, 767–773. [Google Scholar] [CrossRef] [PubMed]
  12. Berkemeier, L.R.; Winslow, J.W.; Kaplan, D.R.; Nikolics, K.; Goeddel, D.V.; Rosenthal, A. Neurotrophin-5: A Novel Neurotrophic Factor That Activates Trk and trkB. Neuron 1991, 7, 857–866. [Google Scholar] [CrossRef] [PubMed]
  13. Hallböök, F.; Ibáñez, C.F.; Persson, H. Evolutionary Studies of the Nerve Growth Factor Family Reveal a Novel Member Abundantly Expressed in Xenopus Ovary. Neuron 1991, 6, 845–858. [Google Scholar] [CrossRef] [PubMed]
  14. Ip, N.Y.; Ibáñez, C.F.; Nye, S.H.; McClain, J.; Jones, P.F.; Gies, D.R.; Belluscio, L.; Le Beau, M.M.; Espinosa, R.; Squinto, S.P. Mammalian Neurotrophin-4: Structure, Chromosomal Localization, Tissue Distribution, and Receptor Specificity. Proc. Natl. Acad. Sci. USA 1992, 89, 3060–3064. [Google Scholar] [CrossRef] [PubMed]
  15. Timmusk, T.; Belluardo, N.; Metsis, M.; Persson, H. Widespread and Developmentally Regulated Expression of Neurotrophin-4 mRNA in Rat Brain and Peripheral Tissues. Eur. J. Neurosci. 1993, 5, 605–613. [Google Scholar] [CrossRef] [PubMed]
  16. Hofer, M.; Pagliusi, S.R.; Hohn, A.; Leibrock, J.; Barde, Y.A. Regional Distribution of Brain-Derived Neurotrophic Factor mRNA in the Adult Mouse Brain. EMBO J. 1990, 9, 2459–2464. [Google Scholar] [CrossRef] [PubMed]
  17. Esvald, E.-E.; Tuvikene, J.; Kiir, C.S.; Avarlaid, A.; Tamberg, L.; Sirp, A.; Shubina, A.; Cabrera-Cabrera, F.; Pihlak, A.; Koppel, I.; et al. Revisiting the Expression of BDNF and Its Receptors in Mammalian Development. Front. Mol. Neurosci. 2023, 16, 1182499. [Google Scholar] [CrossRef] [PubMed]
  18. Murase, K.; Igarashi, K.; Hayashi, K. Neurotrophin-3 (NT-3) Levels in the Developing Rat Nervous System and in Human Samples. Clin. Chim. Acta 1994, 227, 23–36. [Google Scholar] [CrossRef] [PubMed]
  19. Nishio, T.; Akiguchi, I.; Furukawa, S. Detailed Distribution of Nerve Growth Factor in Rat Brain Determined by a Highly Sensitive Enzyme Immunoassay. Exp. Neurol. 1992, 116, 76–84. [Google Scholar] [CrossRef]
  20. Conner, J.M.; Lauterborn, J.C.; Yan, Q.; Gall, C.M.; Varon, S. Distribution of Brain-Derived Neurotrophic Factor (BDNF) Protein and mRNA in the Normal Adult Rat CNS: Evidence for Anterograde Axonal Transport. J. Neurosci. 1997, 17, 2295–2313. [Google Scholar] [CrossRef]
  21. Friedman, W.J.; Black, I.B.; Kaplan, D.R. Distribution of the Neurotrophins Brain-Derived Neurotrophic Factor, Neurotrophin-3, and Neurotrophin-4/5 in the Postnatal Rat Brain: An Immunocytochemical Study. Neuroscience 1998, 84, 101–114. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, H.-T.; Li, L.-Y.; Zou, X.-L.; Song, X.-B.; Hu, Y.-L.; Feng, Z.-T.; Wang, T.T.-H. Immunohistochemical Distribution of NGF, BDNF, NT-3, and NT-4 in Adult Rhesus Monkey Brains. J. Histochem. Cytochem. 2007, 55, 1–19. [Google Scholar] [CrossRef] [PubMed]
  23. Arévalo, J.C.; Deogracias, R. Mechanisms Controlling the Expression and Secretion of BDNF. Biomolecules 2023, 13, 789. [Google Scholar] [CrossRef] [PubMed]
  24. Timmusk, T.; Palm, K.; Metsis, M.; Reintam, T.; Paalme, V.; Saarma, M.; Persson, H. Multiple Promoters Direct Tissue-Specific Expression of the Rat BDNF Gene. Neuron 1993, 10, 475–489. [Google Scholar] [CrossRef] [PubMed]
  25. You, H.; Lu, B. Diverse Functions of Multiple Bdnf Transcripts Driven by Distinct Bdnf Promoters. Biomolecules 2023, 13, 655. [Google Scholar] [CrossRef] [PubMed]
  26. Colliva, A.; Tongiorgi, E. Distinct Role of 5′UTR Sequences in Dendritic Trafficking of BDNF mRNA: Additional Mechanisms for the BDNF Splice Variants Spatial Code. Mol. Brain 2021, 14, 10. [Google Scholar] [CrossRef] [PubMed]
  27. Maynard, K.R.; Hobbs, J.W.; Sukumar, M.; Kardian, A.S.; Jimenez, D.V.; Schloesser, R.J.; Martinowich, K. Bdnf mRNA Splice Variants Differentially Impact CA1 and CA3 Dendrite Complexity and Spine Morphology in the Hippocampus. Brain Struct. Funct. 2017, 222, 3295–3307. [Google Scholar] [CrossRef] [PubMed]
  28. You, H.; Chu, P.; Guo, W.; Lu, B. A Subpopulation of Bdnf-E1–Expressing Glutamatergic Neurons in the Lateral Hypothalamus Critical for Thermogenesis Control. Mol. Metab. 2020, 31, 109–123. [Google Scholar] [CrossRef] [PubMed]
  29. Maynard, K.R.; Hill, J.L.; Calcaterra, N.E.; Palko, M.E.; Kardian, A.; Paredes, D.; Sukumar, M.; Adler, B.D.; Jimenez, D.V.; Schloesser, R.J.; et al. Functional Role of BDNF Production from Unique Promoters in Aggression and Serotonin Signaling. Neuropsychopharmacology 2016, 41, 1943–1955. [Google Scholar] [CrossRef]
  30. Maynard, K.R.; Hobbs, J.W.; Phan, B.N.; Gupta, A.; Rajpurohit, S.; Williams, C.; Rajpurohit, A.; Shin, J.H.; Jaffe, A.E.; Martinowich, K. BDNF-TrkB Signaling in Oxytocin Neurons Contributes to Maternal Behavior. eLife 2018, 7, e33676. [Google Scholar] [CrossRef]
  31. Hong, E.J.; McCord, A.E.; Greenberg, M.E. A Biological Function for the Neuronal Activity-Dependent Component of Bdnf Transcription in the Development of Cortical Inhibition. Neuron 2008, 60, 610–624. [Google Scholar] [CrossRef] [PubMed]
  32. Sakata, K.; Woo, N.H.; Martinowich, K.; Greene, J.S.; Schloesser, R.J.; Shen, L.; Lu, B. Critical Role of Promoter IV-Driven BDNF Transcription in GABAergic Transmission and Synaptic Plasticity in the Prefrontal Cortex. Proc. Natl. Acad. Sci. USA 2009, 106, 5942–5947. [Google Scholar] [CrossRef] [PubMed]
  33. Kolbeck, R.; Bartke, I.; Eberle, W.; Barde, Y. Brain-Derived Neurotrophic Factor Levels in the Nervous System of Wild-Type and Neurotrophin Gene Mutant Mice. J. Neurochem. 1999, 72, 1930–1938. [Google Scholar] [CrossRef] [PubMed]
  34. Barbacid, M. The Trk Family of Neurotrophin Receptors. J. Neurobiol. 1994, 25, 1386–1403. [Google Scholar] [CrossRef] [PubMed]
  35. Huang, E.J.; Reichardt, L.F. Trk Receptors: Roles in Neuronal Signal Transduction. Annu. Rev. Biochem. 2003, 72, 609–642. [Google Scholar] [CrossRef] [PubMed]
  36. Rodriguez-Tebar, A.; Dechant, G.; Barde, Y.-A. Binding of Brain-Derived Neurotrophic Factor to the Nerve Growth Factor Receptor. Neuron 1990, 4, 487–492. [Google Scholar] [CrossRef] [PubMed]
  37. Rodríguez-Tébar, A.; Dechant, G.; Götz, R.; Barde, Y.A. Binding of Neurotrophin-3 to Its Neuronal Receptors and Interactions with Nerve Growth Factor and Brain-Derived Neurotrophic Factor. EMBO J. 1992, 11, 917–922. [Google Scholar] [CrossRef] [PubMed]
  38. Rydén, M.; Murray-Rust, J.; Glass, D.; Ilag, L.L.; Trupp, M.; Yancopoulos, G.D.; McDonald, N.Q.; Ibáñez, C.F. Functional Analysis of Mutant Neurotrophins Deficient in Low-Affinity Binding Reveals a Role for p75LNGFR in NT-4 Signalling. EMBO J. 1995, 14, 1979–1990. [Google Scholar] [CrossRef] [PubMed]
  39. Almeida, R.D.; Duarte, C.B. p75NTR Processing and Signaling: Functional Role. In Handbook of Neurotoxicity; Kostrzewa, R.M., Ed.; Springer: New York, NY, USA, 2014; pp. 1899–1923. ISBN 978-1-4614-5835-7. [Google Scholar]
  40. Conroy, J.N.; Coulson, E.J. High-Affinity TrkA and P75 Neurotrophin Receptor Complexes: A Twisted Affair. J. Biol. Chem. 2022, 298, 101568. [Google Scholar] [CrossRef]
  41. Valenzuela, D.M.; Maisonpierre, P.C.; Glass, D.J.; Rojas, E.; Nuñez, L.; Kong, Y.; Gies, D.R.; Stitt, T.N.; Ip, N.Y.; Yancopoulos, G.D. Alternative Forms of Rat TrkC with Different Functional Capabilities. Neuron 1993, 10, 963–974. [Google Scholar] [CrossRef]
  42. Lee, F.S.; Chao, M.V. Activation of Trk Neurotrophin Receptors in the Absence of Neurotrophins. Proc. Natl. Acad. Sci. USA 2001, 98, 3555–3560. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, F.S.; Rajagopal, R.; Kim, A.H.; Chang, P.C.; Chao, M.V. Activation of Trk Neurotrophin Receptor Signaling by Pituitary Adenylate Cyclase-Activating Polypeptides. J. Biol. Chem. 2002, 277, 9096–9102. [Google Scholar] [CrossRef]
  44. Rajagopal, R.; Chen, Z.-Y.; Lee, F.S.; Chao, M.V. Transactivation of Trk Neurotrophin Receptors by G-Protein-Coupled Receptor Ligands Occurs on Intracellular Membranes. J. Neurosci. 2004, 24, 6650–6658. [Google Scholar] [CrossRef] [PubMed]
  45. Jeanneteau, F.; Garabedian, M.J.; Chao, M.V. Activation of Trk Neurotrophin Receptors by Glucocorticoids Provides a Neuroprotective Effect. Proc. Natl. Acad. Sci. USA 2008, 105, 4862–4867. [Google Scholar] [CrossRef] [PubMed]
  46. Huang, Y.Z.; Pan, E.; Xiong, Z.-Q.; McNamara, J.O. Zinc-Mediated Transactivation of TrkB Potentiates the Hippocampal Mossy Fiber-CA3 Pyramid Synapse. Neuron 2008, 57, 546–558. [Google Scholar] [CrossRef]
  47. Puehringer, D.; Orel, N.; Lüningschrör, P.; Subramanian, N.; Herrmann, T.; Chao, M.V.; Sendtner, M. EGF Transactivation of Trk Receptors Regulates the Migration of Newborn Cortical Neurons. Nat. Neurosci. 2013, 16, 407–415. [Google Scholar] [CrossRef] [PubMed]
  48. Mitre, M.; Saadipour, K.; Williams, K.; Khatri, L.; Froemke, R.C.; Chao, M.V. Transactivation of TrkB Receptors by Oxytocin and Its G Protein-Coupled Receptor. Front. Mol. Neurosci. 2022, 15, 891537. [Google Scholar] [CrossRef] [PubMed]
  49. Snider, W.D. Functions of the Neurotrophins during Nervous System Development: What the Knockouts Are Teaching Us. Cell 1994, 77, 627–638. [Google Scholar] [CrossRef] [PubMed]
  50. Ernfors, P.; Lee, K.-F.; Jaenisch, R. Mice Lacking Brain-Derived Neurotrophic Factor Develop with Sensory Deficits. Nature 1994, 368, 147–150. [Google Scholar] [CrossRef]
  51. Liu, X.; Ernfors, P.; Wu, H.; Jaenisch, R. Sensory but Not Motor Neuron Deficits in Mice Lacking NT4 and BDNF. Nature 1995, 375, 238–241. [Google Scholar] [CrossRef]
  52. Chen, K.S.; Nishimura, M.C.; Armanini, M.P.; Crowley, C.; Spencer, S.D.; Phillips, H.S. Disruption of a Single Allele of the Nerve Growth Factor Gene Results in Atrophy of Basal Forebrain Cholinergic Neurons and Memory Deficits. J. Neurosci. 1997, 17, 7288–7296. [Google Scholar] [CrossRef]
  53. Rauskolb, S.; Zagrebelsky, M.; Dreznjak, A.; Deogracias, R.; Matsumoto, T.; Wiese, S.; Erne, B.; Sendtner, M.; Schaeren-Wiemers, N.; Korte, M.; et al. Global Deprivation of Brain-Derived Neurotrophic Factor in the CNS Reveals an Area-Specific Requirement for Dendritic Growth. J. Neurosci. 2010, 30, 1739–1749. [Google Scholar] [CrossRef]
  54. Reichardt, L.F. Neurotrophin-Regulated Signalling Pathways. Philos. Trans. R. Soc. B Biol. Sci. 2006, 361, 1545–1564. [Google Scholar] [CrossRef]
  55. Park, H.; Poo, M. Neurotrophin Regulation of Neural Circuit Development and Function. Nat. Rev. Neurosci. 2013, 14, 7–23. [Google Scholar] [CrossRef]
  56. Cao, T.; Matyas, J.J.; Renn, C.L.; Faden, A.I.; Dorsey, S.G.; Wu, J. Function and Mechanisms of Truncated BDNF Receptor TrkB.T1 in Neuropathic Pain. Cells 2020, 9, 1194. [Google Scholar] [CrossRef]
  57. Biffo, S.; Offenhäuser, N.; Carter, B.D.; Barde, Y.-A. Selective Binding and Internalisation by Truncated Receptors Restrict the Availability of BDNF during Development. Development 1995, 121, 2461–2470. [Google Scholar] [CrossRef] [PubMed]
  58. Holt, L.M.; Hernandez, R.D.; Pacheco, N.L.; Torres Ceja, B.; Hossain, M.; Olsen, M.L. Astrocyte Morphogenesis Is Dependent on BDNF Signaling via Astrocytic TrkB.T1. eLife 2019, 8, e44667. [Google Scholar] [CrossRef]
  59. Aroeira, R.I.; Sebastião, A.M.; Valente, C.A. BDNF, via Truncated TrkB Receptor, Modulates GlyT1 and GlyT2 in Astrocytes: Modulation of GlyT by BDNF. Glia 2015, 63, 2181–2197. [Google Scholar] [CrossRef] [PubMed]
  60. Dorsey, S.G.; Renn, C.L.; Carim-Todd, L.; Barrick, C.A.; Bambrick, L.; Krueger, B.K.; Ward, C.W.; Tessarollo, L. In Vivo Restoration of Physiological Levels of Truncated TrkB.T1 Receptor Rescues Neuronal Cell Death in a Trisomic Mouse Model. Neuron 2006, 51, 21–28. [Google Scholar] [CrossRef] [PubMed]
  61. Saba, J.; Turati, J.; Ramírez, D.; Carniglia, L.; Durand, D.; Lasaga, M.; Caruso, C. Astrocyte Truncated Tropomyosin Receptor Kinase B Mediates Brain-derived Neurotrophic Factor Anti-apoptotic Effect Leading to Neuroprotection. J. Neurochem. 2018, 146, 686–702. [Google Scholar] [CrossRef]
  62. Michaelsen, K.; Zagrebelsky, M.; Berndt-Huch, J.; Polack, M.; Buschler, A.; Sendtner, M.; Korte, M. Neurotrophin Receptors TrkB.T1 and P75 NTR Cooperate in Modulating Both Functional and Structural Plasticity in Mature Hippocampal Neurons. Eur. J. Neurosci. 2010, 32, 1854–1865. [Google Scholar] [CrossRef] [PubMed]
  63. Underwood, C.K.; Coulson, E.J. The P75 Neurotrophin Receptor. Int. J. Biochem. Cell Biol. 2008, 40, 1664–1668. [Google Scholar] [CrossRef] [PubMed]
  64. Fahnestock, M.; Michalski, B.; Xu, B.; Coughlin, M.D. The Precursor Pro-Nerve Growth Factor Is the Predominant Form of Nerve Growth Factor in Brain and Is Increased in Alzheimer’s Disease. Mol. Cell. Neurosci. 2001, 18, 210–220. [Google Scholar] [CrossRef] [PubMed]
  65. Bruno, M.A.; Leon, W.C.; Fragoso, G.; Mushynski, W.E.; Almazan, G.; Cuello, A.C. Amyloid β-Induced Nerve Growth Factor Dysmetabolism in Alzheimer Disease. J. Neuropathol. Exp. Neurol. 2009, 68, 857–869. [Google Scholar] [CrossRef] [PubMed]
  66. Yang, J.; Harte-Hargrove, L.C.; Siao, C.-J.; Marinic, T.; Clarke, R.; Ma, Q.; Jing, D.; LaFrancois, J.J.; Bath, K.G.; Mark, W.; et al. proBDNF Negatively Regulates Neuronal Remodeling, Synaptic Transmission, and Synaptic Plasticity in Hippocampus. Cell Rep. 2014, 7, 796–806. [Google Scholar] [CrossRef] [PubMed]
  67. Aloyz, R.S.; Bamji, S.X.; Pozniak, C.D.; Toma, J.G.; Atwal, J.; Kaplan, D.R.; Miller, F.D. P53 Is Essential for Developmental Neuron Death as Regulated by the TrkA and P75 Neurotrophin Receptors. J. Cell Biol. 1998, 143, 1691–1703. [Google Scholar] [CrossRef] [PubMed]
  68. Linggi, M.S.; Burke, T.L.; Williams, B.B.; Harrington, A.; Kraemer, R.; Hempstead, B.L.; Yoon, S.O.; Carter, B.D. Neurotrophin Receptor Interacting Factor (NRIF) Is an Essential Mediator of Apoptotic Signaling by the P75 Neurotrophin Receptor. J. Biol. Chem. 2005, 280, 13801–13808. [Google Scholar] [CrossRef] [PubMed]
  69. Becker, E.B.E.; Howell, J.; Kodama, Y.; Barker, P.A.; Bonni, A. Characterization of the C-Jun N-Terminal Kinase-Bim EL Signaling Pathway in Neuronal Apoptosis. J. Neurosci. 2004, 24, 8762–8770. [Google Scholar] [CrossRef] [PubMed]
  70. Bhakar, A.L.; Howell, J.L.; Paul, C.E.; Salehi, A.H.; Becker, E.B.E.; Said, F.; Bonni, A.; Barker, P.A. Apoptosis Induced by p75NTR Overexpression Requires Jun Kinase-Dependent Phosphorylation of Bad. J. Neurosci. 2003, 23, 11373–11381. [Google Scholar] [CrossRef] [PubMed]
  71. Yamashita, T.; Tucker, K.L.; Barde, Y.-A. Neurotrophin Binding to the P75 Receptor Modulates Rho Activity and Axonal Outgrowth. Neuron 1999, 24, 585–593. [Google Scholar] [CrossRef]
  72. Shonukan, T.; Bagayogo, I.; McCrea, P.; Chao, M.; Hempstead, B. Neurotrophin-Induced Melanoma Cell Migration Is Mediated through the Actin-Bundling Protein Fascin. Oncogene 2003, 22, 3616–3623. [Google Scholar] [CrossRef]
  73. Zagrebelsky, M.; Holz, A.; Dechant, G.; Barde, Y.-A.; Bonhoeffer, T.; Korte, M. The P75 Neurotrophin Receptor Negatively Modulates Dendrite Complexity and Spine Density in Hippocampal Neurons. J. Neurosci. 2005, 25, 9989–9999. [Google Scholar] [CrossRef]
  74. Singh, K.K.; Park, K.J.; Hong, E.J.; Kramer, B.M.; Greenberg, M.E.; Kaplan, D.R.; Miller, F.D. Developmental Axon Pruning Mediated by BDNF-p75NTR–Dependent Axon Degeneration. Nat. Neurosci. 2008, 11, 649–658. [Google Scholar] [CrossRef]
  75. Deinhardt, K.; Kim, T.; Spellman, D.S.; Mains, R.E.; Eipper, B.A.; Neubert, T.A.; Chao, M.V.; Hempstead, B.L. Neuronal Growth Cone Retraction Relies on Proneurotrophin Receptor Signaling Through Rac. Sci. Signal. 2011, 4, ra82. [Google Scholar] [CrossRef] [PubMed]
  76. Matsumoto, T.; Rauskolb, S.; Polack, M.; Klose, J.; Kolbeck, R.; Korte, M.; Barde, Y.-A. Biosynthesis and Processing of Endogenous BDNF: CNS Neurons Store and Secrete BDNF, not pro-BDNF. Nat. Neurosci. 2008, 11, 131–133. [Google Scholar] [CrossRef] [PubMed]
  77. Nagappan, G.; Zaitsev, E.; Senatorov, V.V.; Yang, J.; Hempstead, B.L.; Lu, B. Control of Extracellular Cleavage of ProBDNF by High Frequency Neuronal Activity. Proc. Natl. Acad. Sci. USA 2009, 106, 1267–1272. [Google Scholar] [CrossRef]
  78. Woo, N.H.; Teng, H.K.; Siao, C.-J.; Chiaruttini, C.; Pang, P.T.; Milner, T.A.; Hempstead, B.L.; Lu, B. Activation of p75NTR by proBDNF Facilitates Hippocampal Long-Term Depression. Nat. Neurosci. 2005, 8, 1069–1077. [Google Scholar] [CrossRef] [PubMed]
  79. Frade, J.M.; López-Sánchez, N. A Novel Hypothesis for Alzheimer Disease Based on Neuronal Tetraploidy Induced by P75 NTR. Cell Cycle 2010, 9, 1934–1941. [Google Scholar] [CrossRef]
  80. Frade, J.M.; Ovejero-Benito, M.C. Neuronal Cell Cycle: The Neuron Itself and Its Circumstances. Cell Cycle 2015, 14, 712–720. [Google Scholar] [CrossRef]
  81. Camuso, S.; La Rosa, P.; Fiorenza, M.T.; Canterini, S. Pleiotropic Effects of BDNF on the Cerebellum and Hippocampus: Implications for Neurodevelopmental Disorders. Neurobiol. Dis. 2022, 163, 105606. [Google Scholar] [CrossRef]
  82. Vicario-Abejón, C.; Collin, C.; McKay, R.D.G.; Segal, M. Neurotrophins Induce Formation of Functional Excitatory and Inhibitory Synapses between Cultured Hippocampal Neurons. J. Neurosci. 1998, 18, 7256–7271. [Google Scholar] [CrossRef] [PubMed]
  83. Waterhouse, E.G.; An, J.J.; Orefice, L.L.; Baydyuk, M.; Liao, G.-Y.; Zheng, K.; Lu, B.; Xu, B. BDNF Promotes Differentiation and Maturation of Adult-Born Neurons through GABAergic Transmission. J. Neurosci. 2012, 32, 14318–14330. [Google Scholar] [CrossRef] [PubMed]
  84. Zhu, G.; Sun, X.; Yang, Y.; Du, Y.; Lin, Y.; Xiang, J.; Zhou, N. Reduction of BDNF Results in GABAergic Neuroplasticity Dysfunction and Contributes to Late-Life Anxiety Disorder. Behav. Neurosci. 2019, 133, 212–224. [Google Scholar] [CrossRef] [PubMed]
  85. Götz, M.; Huttner, W.B. The Cell Biology of Neurogenesis. Nat. Rev. Mol. Cell Biol. 2005, 6, 777–788. [Google Scholar] [CrossRef] [PubMed]
  86. Ortega, J.A.; Alcántara, S. BDNF/MAPK/ERK–Induced BMP7 Expression in the Developing Cerebral Cortex Induces Premature Radial Glia Differentiation and Impairs Neuronal Migration. Cereb. Cortex 2010, 20, 2132–2144. [Google Scholar] [CrossRef] [PubMed]
  87. Lim, L.; Mi, D.; Llorca, A.; Marín, O. Development and Functional Diversification of Cortical Interneurons. Neuron 2018, 100, 294–313. [Google Scholar] [CrossRef] [PubMed]
  88. Llorca, A.; Deogracias, R. Origin, Development, and Synaptogenesis of Cortical Interneurons. Front. Neurosci. 2022, 16, 929469. [Google Scholar] [CrossRef] [PubMed]
  89. Kohara, K.; Kitamura, A.; Adachi, N.; Nishida, M.; Itami, C.; Nakamura, S.; Tsumoto, T. Inhibitory but Not Excitatory Cortical Neurons Require Presynaptic Brain-Derived Neurotrophic Factor for Dendritic Development, as Revealed by Chimera Cell Culture. J. Neurosci. 2003, 23, 6123–6131. [Google Scholar] [CrossRef] [PubMed]
  90. Gorski, J.A.; Zeiler, S.R.; Tamowski, S.; Jones, K.R. Brain-Derived Neurotrophic Factor Is Required for the Maintenance of Cortical Dendrites. J. Neurosci. 2003, 23, 6856–6865. [Google Scholar] [CrossRef]
  91. Patz, S. Parvalbumin Expression in Visual Cortical Interneurons Depends on Neuronal Activity and TrkB Ligands during an Early Period of Postnatal Development. Cereb. Cortex 2004, 14, 342–351. [Google Scholar] [CrossRef]
  92. Du, Z.; Tertrais, M.; Courtand, G.; Leste-Lasserre, T.; Cardoit, L.; Masmejean, F.; Halgand, C.; Cho, Y.H.; Garret, M. Differential Alteration in Expression of Striatal GABAAR Subunits in Mouse Models of Huntington’s Disease. Front. Mol. Neurosci. 2017, 10, 198. [Google Scholar] [CrossRef] [PubMed]
  93. Sánchez-Huertas, C.; Rico, B. CREB-Dependent Regulation of GAD65 Transcription by BDNF/TrkB in Cortical Interneurons. Cereb. Cortex 2011, 21, 777–788. [Google Scholar] [CrossRef] [PubMed]
  94. Holter, M.C.; Hewitt, L.T.; Nishimura, K.J.; Knowles, S.J.; Bjorklund, G.R.; Shah, S.; Fry, N.R.; Rees, K.P.; Gupta, T.A.; Daniels, C.W.; et al. Hyperactive MEK1 Signaling in Cortical GABAergic Neurons Promotes Embryonic Parvalbumin Neuron Loss and Defects in Behavioral Inhibition. Cereb. Cortex 2021, 31, 3064–3081. [Google Scholar] [CrossRef] [PubMed]
  95. Lau, C.G.; Zhang, H.; Murthy, V.N. Deletion of TrkB in Parvalbumin Interneurons Alters Cortical Neural Dynamics. J. Cell. Physiol. 2022, 237, 949–964. [Google Scholar] [CrossRef]
  96. Grabert, J.; Wahle, P. Neuronal Activity and TrkB Ligands Influence Kv3.1b and Kv3.2 Expression in Developing Cortical Interneurons. Neuroscience 2008, 156, 618–629. [Google Scholar] [CrossRef] [PubMed]
  97. Fawcett, J.W.; Oohashi, T.; Pizzorusso, T. The Roles of Perineuronal Nets and the Perinodal Extracellular Matrix in Neuronal Function. Nat. Rev. Neurosci. 2019, 20, 451–465. [Google Scholar] [CrossRef] [PubMed]
  98. Willis, A.; Pratt, J.A.; Morris, B.J. BDNF and JNK Signaling Modulate Cortical Interneuron and Perineuronal Net Development: Implications for Schizophrenia-Linked 16p11.2 Duplication Syndrome. Schizophr. Bull. 2021, 47, 812–826. [Google Scholar] [CrossRef] [PubMed]
  99. Meier, S.; Alfonsi, F.; Kurniawan, N.D.; Milne, M.R.; Kasherman, M.A.; Delogu, A.; Piper, M.; Coulson, E.J. The P75 Neurotrophin Receptor Is Required for the Survival of Neuronal Progenitors and Normal Formation of the Basal Forebrain, Striatum, Thalamus and Neocortex. Development 2019, 146, dev181933. [Google Scholar] [CrossRef]
  100. Baho, E.; Chattopadhyaya, B.; Lavertu-Jolin, M.; Mazziotti, R.; Awad, P.N.; Chehrazi, P.; Groleau, M.; Jahannault-Talignani, C.; Vaucher, E.; Ango, F.; et al. P75 Neurotrophin Receptor Activation Regulates the Timing of the Maturation of Cortical Parvalbumin Interneuron Connectivity and Promotes Juvenile-like Plasticity in Adult Visual Cortex. J. Neurosci. 2019, 39, 4489–4510. [Google Scholar] [CrossRef]
  101. Chehrazi, P.; Lee, K.K.Y.; Lavertu-Jolin, M.; Abbasnejad, Z.; Carreño-Muñoz, M.I.; Chattopadhyaya, B.; Di Cristo, G. The P75 Neurotrophin Receptor in Preadolescent Prefrontal Parvalbumin Interneurons Promotes Cognitive Flexibility in Adult Mice. Biol. Psychiatry 2023, 94, 310–321. [Google Scholar] [CrossRef]
  102. Knowles, R.; Dehorter, N.; Ellender, T. From Progenitors to Progeny: Shaping Striatal Circuit Development and Function. J. Neurosci. 2021, 41, 9483–9502. [Google Scholar] [CrossRef] [PubMed]
  103. Dudman, J.T.; Gerfen, C.R. The Basal Ganglia. In The Rat Nervous System; Elsevier: Amsterdam, The Netherlands, 2015; pp. 391–440. ISBN 978-0-12-374245-2. [Google Scholar]
  104. Chuhma, N.; Tanaka, K.F.; Hen, R.; Rayport, S. Functional Connectome of the Striatal Medium Spiny Neuron. J. Neurosci. 2011, 31, 1183–1192. [Google Scholar] [CrossRef] [PubMed]
  105. Gagnon, D.; Petryszyn, S.; Sanchez, M.G.; Bories, C.; Beaulieu, J.M.; De Koninck, Y.; Parent, A.; Parent, M. Striatal Neurons Expressing D1 and D2 Receptors Are Morphologically Distinct and Differently Affected by Dopamine Denervation in Mice. Sci. Rep. 2017, 7, 41432. [Google Scholar] [CrossRef] [PubMed]
  106. Gerfen, C.R. Segregation of D1 and D2 Dopamine Receptors in the Striatal Direct and Indirect Pathways: An Historical Perspective. Front. Synaptic Neurosci. 2023, 14, 1002960. [Google Scholar] [CrossRef] [PubMed]
  107. Florio, T.M.; Scarnati, E.; Rosa, I.; Di Censo, D.; Ranieri, B.; Cimini, A.; Galante, A.; Alecci, M. The Basal Ganglia: More than Just a Switching Device. CNS Neurosci. Ther. 2018, 24, 677–684. [Google Scholar] [CrossRef] [PubMed]
  108. Eisinger, R.S.; Urdaneta, M.E.; Foote, K.D.; Okun, M.S.; Gunduz, A. Non-Motor Characterization of the Basal Ganglia: Evidence from Human and Non-Human Primate Electrophysiology. Front. Neurosci. 2018, 12, 385. [Google Scholar] [CrossRef] [PubMed]
  109. Baydyuk, M.; Xu, B. BDNF Signaling and Survival of Striatal Neurons. Front. Cell. Neurosci. 2014, 8, 254. [Google Scholar] [CrossRef] [PubMed]
  110. Ehinger, Y.; Soneja, D.; Phamluong, K.; Salvi, A.; Ron, D. Identification of Novel BDNF-Specific Corticostriatal Circuitries. Eneuro 2023, 10, ENEURO.0238-21.2023. [Google Scholar] [CrossRef]
  111. McCutcheon, R.A.; Abi-Dargham, A.; Howes, O.D. Schizophrenia, Dopamine and the Striatum: From Biology to Symptoms. Trends Neurosci. 2019, 42, 205–220. [Google Scholar] [CrossRef]
  112. Ayon-Olivas, M.; Wolf, D.; Andreska, T.; Granado, N.; Lüningschrör, P.; Ip, C.W.; Moratalla, R.; Sendtner, M. Dopaminergic Input Regulates the Sensitivity of Indirect Pathway Striatal Spiny Neurons to Brain-Derived Neurotrophic Factor. Biology 2023, 12, 1360. [Google Scholar] [CrossRef]
  113. Andreska, T.; Lüningschrör, P.; Wolf, D.; McFleder, R.L.; Ayon-Olivas, M.; Rattka, M.; Drechsler, C.; Perschin, V.; Blum, R.; Aufmkolk, S.; et al. DRD1 Signaling Modulates TrkB Turnover and BDNF Sensitivity in Direct Pathway Striatal Medium Spiny Neurons. Cell Rep. 2023, 42, 112575. [Google Scholar] [CrossRef] [PubMed]
  114. Li, Y.; Yui, D.; Luikart, B.W.; McKay, R.M.; Li, Y.; Rubenstein, J.L.; Parada, L.F. Conditional Ablation of Brain-Derived Neurotrophic Factor-TrkB Signaling Impairs Striatal Neuron Development. Proc. Natl. Acad. Sci. USA 2012, 109, 15491–15496. [Google Scholar] [CrossRef] [PubMed]
  115. Baydyuk, M.; Russell, T.; Liao, G.-Y.; Zang, K.; An, J.J.; Reichardt, L.F.; Xu, B. TrkB Receptor Controls Striatal Formation by Regulating the Number of Newborn Striatal Neurons. Proc. Natl. Acad. Sci. USA 2011, 108, 1669–1674. [Google Scholar] [CrossRef] [PubMed]
  116. Baydyuk, M.; Xie, Y.; Tessarollo, L.; Xu, B. Midbrain-Derived Neurotrophins Support Survival of Immature Striatal Projection Neurons. J. Neurosci. 2013, 33, 3363–3369. [Google Scholar] [CrossRef] [PubMed]
  117. Koshimizu, H.; Matsuoka, H.; Nakajima, Y.; Kawai, A.; Ono, J.; Ohta, K.; Miki, T.; Sunagawa, M.; Adachi, N.; Suzuki, S. Brain-derived Neurotrophic Factor Predominantly Regulates the Expression of Synapse-related Genes in the Striatum: Insights from in Vitro Transcriptomics. Neuropsychopharmacol. Rep. 2021, 41, 485–495. [Google Scholar] [CrossRef] [PubMed]
  118. Goggi, J.; Pullar, I.A.; Carney, S.L.; Bradford, H.F. Signalling Pathways Involved in the Short-Term Potentiation of Dopamine Release by BDNF. Brain Res. 2003, 968, 156–161. [Google Scholar] [CrossRef] [PubMed]
  119. Kang, D.-S.; Kim, I.S.; Baik, J.-H.; Kim, D.; Cocco, L.; Suh, P.-G. The Function of PLCγ1 in Developing Mouse mDA System. Adv. Biol. Regul. 2020, 75, 100654. [Google Scholar] [CrossRef] [PubMed]
  120. Kim, H.Y.; Lee, J.; Kim, H.-J.; Lee, B.E.; Jeong, J.; Cho, E.J.; Jang, H.-J.; Shin, K.J.; Kim, M.J.; Chae, Y.C.; et al. PLCγ1 in Dopamine Neurons Critically Regulates Striatal Dopamine Release via VMAT2 and Synapsin III. Exp. Mol. Med. 2023, 55, 2357–2375. [Google Scholar] [CrossRef] [PubMed]
  121. Hourez, R.; Azdad, K.; Vanwalleghem, G.; Roussel, C.; Gall, D.; Schiffmann, S.N. Activation of Protein Kinase C and Inositol 1,4,5-Triphosphate Receptors Antagonistically Modulate Voltage-Gated Sodium Channels in Striatal Neurons. Brain Res. 2005, 1059, 189–196. [Google Scholar] [CrossRef]
  122. Clements, M.A.; Swapna, I.; Morikawa, H. Inositol 1,4,5-Triphosphate Drives Glutamatergic and Cholinergic Inhibition Selectively in Spiny Projection Neurons in the Striatum. J. Neurosci. 2013, 33, 2697–2708. [Google Scholar] [CrossRef]
  123. Kärkkäinen, E.; Yavich, L.; Miettinen, P.O.; Tanila, H. Opposing Effects of APP/PS1 and TrkB.T1 Genotypes on Midbrain Dopamine Neurons and Stimulated Dopamine Release in Vivo. Brain Res. 2015, 1622, 452–465. [Google Scholar] [CrossRef] [PubMed]
  124. Torres-Cruz, F.M.; César Vivar-Cortés, I.; Moran, I.; Mendoza, E.; Gómez-Pineda, V.; García-Sierra, F.; Hernández-Echeagaray, E. NT-4/5 Antagonizes the BDNF Modulation of Corticostriatal Transmission: Role of the TrkB.T1 Receptor. CNS Neurosci. Ther. 2019, 25, 621–631. [Google Scholar] [CrossRef] [PubMed]
  125. Gomes, J.R.; Costa, J.T.; Melo, C.V.; Felizzi, F.; Monteiro, P.; Pinto, M.J.; Inácio, A.R.; Wieloch, T.; Almeida, R.D.; Grãos, M.; et al. Excitotoxicity Downregulates TrkB.FL Signaling and Upregulates the Neuroprotective Truncated TrkB Receptors in Cultured Hippocampal and Striatal Neurons. J. Neurosci. 2012, 32, 4610–4622. [Google Scholar] [CrossRef] [PubMed]
  126. Plotkin, J.L.; Surmeier, D.J. Impaired Striatal Function in Huntington’s Disease Is Due to Aberrant p75NTR Signaling. Rare Dis. 2014, 2, e968482. [Google Scholar] [CrossRef] [PubMed]
  127. Wehner, A.B.; Milen, A.M.; Albin, R.L.; Pierchala, B.A. The P75 Neurotrophin Receptor Augments Survival Signaling in the Striatum of Pre-Symptomatic Q175 Mice. Neuroscience 2016, 324, 297–306. [Google Scholar] [CrossRef] [PubMed]
  128. Liu, Z.; Yan, A.; Zhao, J.; Yang, S.; Song, L.; Liu, Z. The P75 Neurotrophin Receptor as a Novel Intermediate in L-Dopa-Induced Dyskinesia in Experimental Parkinson’s Disease. Exp. Neurol. 2021, 342, 113740. [Google Scholar] [CrossRef] [PubMed]
  129. Darcq, E.; Morisot, N.; Phamluong, K.; Warnault, V.; Jeanblanc, J.; Longo, F.M.; Massa, S.M.; Ron, D. The Neurotrophic Factor Receptor P75 in the Rat Dorsolateral Striatum Drives Excessive Alcohol Drinking. J. Neurosci. 2016, 36, 10116–10127. [Google Scholar] [CrossRef] [PubMed]
  130. Ramamoorthi, K.; Lin, Y. The Contribution of GABAergic Dysfunction to Neurodevelopmental Disorders. Trends Mol. Med. 2011, 17, 452–462. [Google Scholar] [CrossRef]
  131. Jiang, C.-C.; Lin, L.-S.; Long, S.; Ke, X.-Y.; Fukunaga, K.; Lu, Y.-M.; Han, F. Signalling Pathways in Autism Spectrum Disorder: Mechanisms and Therapeutic Implications. Signal Transduct. Target. Ther. 2022, 7, 229. [Google Scholar] [CrossRef]
  132. Bicks, L.K.; Geschwind, D.H. Functional Neurogenomics in Autism Spectrum Disorders: A Decade of Progress. Curr. Opin. Neurobiol. 2024, 86, 102858. [Google Scholar] [CrossRef]
  133. Inestrosa, N.C.; Varela-Nallar, L. Wnt Signalling in Neuronal Differentiation and Development. Cell Tissue Res. 2015, 359, 215–223. [Google Scholar] [CrossRef] [PubMed]
  134. Chen, Y.; Huang, W.-C.; Sejourne, J.; Clipperton-Allen, A.E.; Page, D.T. Pten Mutations Alter Brain Growth Trajectory and Allocation of Cell Types through Elevated -Catenin Signaling. J. Neurosci. 2015, 35, 10252–10267. [Google Scholar] [CrossRef] [PubMed]
  135. Lu, B. BDNF and Activity-Dependent Synaptic Modulation: Figure 1. Learn. Mem. 2003, 10, 86–98. [Google Scholar] [CrossRef] [PubMed]
  136. West, A.E.; Chen, W.G.; Dalva, M.B.; Dolmetsch, R.E.; Kornhauser, J.M.; Shaywitz, A.J.; Takasu, M.A.; Tao, X.; Greenberg, M.E. Calcium Regulation of Neuronal Gene Expression. Proc. Natl. Acad. Sci. USA 2001, 98, 11024–11031. [Google Scholar] [CrossRef] [PubMed]
  137. Ilchibaeva, T.; Tsybko, A.; Lipnitskaya, M.; Eremin, D.; Milutinovich, K.; Naumenko, V.; Popova, N. Brain-Derived Neurotrophic Factor (BDNF) in Mechanisms of Autistic-like Behavior in BTBR Mice: Crosstalk with the Dopaminergic Brain System. Biomedicines 2023, 11, 1482. [Google Scholar] [CrossRef] [PubMed]
  138. Sgritta, M.; Vignoli, B.; Pimpinella, D.; Griguoli, M.; Santi, S.; Bialowas, A.; Wiera, G.; Zacchi, P.; Malerba, F.; Marchetti, C.; et al. Impaired Synaptic Plasticity in an Animal Model of Autism Exhibiting Early Hippocampal GABAergic-BDNF/TrkB Signaling Alterations. iScience 2023, 26, 105728. [Google Scholar] [CrossRef] [PubMed]
  139. Bludau, A.; Royer, M.; Meister, G.; Neumann, I.D.; Menon, R. Epigenetic Regulation of the Social Brain. Trends Neurosci. 2019, 42, 471–484. [Google Scholar] [CrossRef]
  140. Zhubi, A.; Chen, Y.; Dong, E.; Cook, E.H.; Guidotti, A.; Grayson, D.R. Increased Binding of MeCP2 to the GAD1 and RELN Promoters May Be Mediated by an Enrichment of 5-hmC in Autism Spectrum Disorder (ASD) Cerebellum. Transl. Psychiatry 2014, 4, e349. [Google Scholar] [CrossRef]
  141. Egan, M.F.; Kojima, M.; Callicott, J.H.; Goldberg, T.E.; Kolachana, B.S.; Bertolino, A.; Zaitsev, E.; Gold, B.; Goldman, D.; Dean, M.; et al. The BDNF Val66met Polymorphism Affects Activity-Dependent Secretion of BDNF and Human Memory and Hippocampal Function. Cell 2003, 112, 257–269. [Google Scholar] [CrossRef]
  142. Uegaki, K.; Kumanogoh, H.; Mizui, T.; Hirokawa, T.; Ishikawa, Y.; Kojima, M. BDNF Binds Its Pro-Peptide with High Affinity and the Common Val66Met Polymorphism Attenuates the Interaction. Int. J. Mol. Sci. 2017, 18, 1042. [Google Scholar] [CrossRef]
  143. Ma, K.; Taylor, C.; Williamson, M.; Newton, S.S.; Qin, L. Diminished Activity-Dependent BDNF Signaling Differentially Causes Autism-like Behavioral Deficits in Male and Female Mice. Front. Psychiatry 2023, 14, 1182472. [Google Scholar] [CrossRef] [PubMed]
  144. Neul, J.L.; Kaufmann, W.E.; Glaze, D.G.; Christodoulou, J.; Clarke, A.J.; Bahi-Buisson, N.; Leonard, H.; Bailey, M.E.S.; Schanen, N.C.; Zappella, M.; et al. Rett Syndrome: Revised Diagnostic Criteria and Nomenclature. Ann. Neurol. 2010, 68, 944–950. [Google Scholar] [CrossRef] [PubMed]
  145. Li, W.; Pozzo-Miller, L. BDNF Deregulation in Rett Syndrome. Neuropharmacology 2014, 76, 737–746. [Google Scholar] [CrossRef] [PubMed]
  146. Ure, K.; Lu, H.; Wang, W.; Ito-Ishida, A.; Wu, Z.; He, L.; Sztainberg, Y.; Chen, W.; Tang, J.; Zoghbi, H.Y. Restoration of Mecp2 Expression in GABAergic Neurons Is Sufficient to Rescue Multiple Disease Features in a Mouse Model of Rett Syndrome. eLife 2016, 5, e14198. [Google Scholar] [CrossRef] [PubMed]
  147. Pejhan, S.; Rastegar, M. Role of DNA Methyl-CpG-Binding Protein MeCP2 in Rett Syndrome Pathobiology and Mechanism of Disease. Biomolecules 2021, 11, 75. [Google Scholar] [CrossRef] [PubMed]
  148. Ebert, D.H.; Gabel, H.W.; Robinson, N.D.; Kastan, N.R.; Hu, L.S.; Cohen, S.; Navarro, A.J.; Lyst, M.J.; Ekiert, R.; Bird, A.P.; et al. Activity-Dependent Phosphorylation of MeCP2 Threonine 308 Regulates Interaction with NCoR. Nature 2013, 499, 341–345. [Google Scholar] [CrossRef] [PubMed]
  149. Tillotson, R.; Bird, A. The Molecular Basis of MeCP2 Function in the Brain. J. Mol. Biol. 2020, 432, 1602–1623. [Google Scholar] [CrossRef] [PubMed]
  150. Chahrour, M.; Jung, S.Y.; Shaw, C.; Zhou, X.; Wong, S.T.C.; Qin, J.; Zoghbi, H.Y. MeCP2, a Key Contributor to Neurological Disease, Activates and Represses Transcription. Science 2008, 320, 1224–1229. [Google Scholar] [CrossRef] [PubMed]
  151. Vuu, Y.M.; Roberts, C.-T.; Rastegar, M. MeCP2 Is an Epigenetic Factor That Links DNA Methylation with Brain Metabolism. Int. J. Mol. Sci. 2023, 24, 4218. [Google Scholar] [CrossRef]
  152. Tzeng, C.P.; Whitwam, T.; Boxer, L.D.; Li, E.; Silberfeld, A.; Trowbridge, S.; Mei, K.; Lin, C.; Shamah, R.; Griffith, E.C.; et al. Activity-Induced MeCP2 Phosphorylation Regulates Retinogeniculate Synapse Refinement. Proc. Natl. Acad. Sci. USA 2023, 120, e2310344120. [Google Scholar] [CrossRef]
  153. Cohen, S.; Gabel, H.W.; Hemberg, M.; Hutchinson, A.N.; Sadacca, L.A.; Ebert, D.H.; Harmin, D.A.; Greenberg, R.S.; Verdine, V.K.; Zhou, Z.; et al. Genome-Wide Activity-Dependent MeCP2 Phosphorylation Regulates Nervous System Development and Function. Neuron 2011, 72, 72–85. [Google Scholar] [CrossRef]
  154. Kim, J.-W.; Autry, A.E.; Na, E.S.; Adachi, M.; Björkholm, C.; Kavalali, E.T.; Monteggia, L.M. Sustained Effects of Rapidly Acting Antidepressants Require BDNF-Dependent MeCP2 Phosphorylation. Nat. Neurosci. 2021, 24, 1100–1109. [Google Scholar] [CrossRef]
  155. Lyst, M.J.; Ekiert, R.; Ebert, D.H.; Merusi, C.; Nowak, J.; Selfridge, J.; Guy, J.; Kastan, N.R.; Robinson, N.D.; De Lima Alves, F.; et al. Rett Syndrome Mutations Abolish the Interaction of MeCP2 with the NCoR/SMRT Co-Repressor. Nat. Neurosci. 2013, 16, 898–902. [Google Scholar] [CrossRef]
  156. Li, W. Excitation and Inhibition Imbalance in Rett Syndrome. Front. Neurosci. 2022, 16, 825063. [Google Scholar] [CrossRef] [PubMed]
  157. Chang, Q.; Khare, G.; Dani, V.; Nelson, S.; Jaenisch, R. The Disease Progression of Mecp2 Mutant Mice Is Affected by the Level of BDNF Expression. Neuron 2006, 49, 341–348. [Google Scholar] [CrossRef] [PubMed]
  158. Larimore, J.L.; Chapleau, C.A.; Kudo, S.; Theibert, A.; Percy, A.K.; Pozzo-Miller, L. Bdnf Overexpression in Hippocampal Neurons Prevents Dendritic Atrophy Caused by Rett-Associated MECP2 Mutations. Neurobiol. Dis. 2009, 34, 199–211. [Google Scholar] [CrossRef]
  159. Massa, S.M.; Yang, T.; Xie, Y.; Shi, J.; Bilgen, M.; Joyce, J.N.; Nehama, D.; Rajadas, J.; Longo, F.M. Small Molecule BDNF Mimetics Activate TrkB Signaling and Prevent Neuronal Degeneration in Rodents. J. Clin. Investig. 2010, 120, 1774–1785. [Google Scholar] [CrossRef]
  160. Johnson, R.A.; Lam, M.; Punzo, A.M.; Li, H.; Lin, B.R.; Ye, K.; Mitchell, G.S.; Chang, Q. 7,8-Dihydroxyflavone Exhibits Therapeutic Efficacy in a Mouse Model of Rett Syndrome. J. Appl. Physiol. 2012, 112, 704–710. [Google Scholar] [CrossRef]
  161. Li, W.; Bellot-Saez, A.; Phillips, M.L.; Yang, T.; Longo, F.M.; Pozzo-Miller, L. A Small-Molecule TrkB Ligand Restores Hippocampal Synaptic Plasticity and Object Location Memory in Rett Syndrome Mice. Dis. Model. Mech. 2017, 10, 837–845. [Google Scholar] [CrossRef]
  162. Deogracias, R.; Yazdani, M.; Dekkers, M.P.J.; Guy, J.; Ionescu, M.C.S.; Vogt, K.E.; Barde, Y.-A. Fingolimod, a Sphingosine-1 Phosphate Receptor Modulator, Increases BDNF Levels and Improves Symptoms of a Mouse Model of Rett Syndrome. Proc. Natl. Acad. Sci. USA 2012, 109, 14230–14235. [Google Scholar] [CrossRef]
  163. Naegelin, Y.; Kuhle, J.; Schädelin, S.; Datta, A.N.; Magon, S.; Amann, M.; Barro, C.; Ramelli, G.P.; Heesom, K.; Barde, Y.-A.; et al. Fingolimod in Children with Rett Syndrome: The FINGORETT Study. Orphanet J. Rare Dis. 2021, 16, 19. [Google Scholar] [CrossRef] [PubMed]
  164. Keam, S.J. Trofinetide: First Approval. Drugs 2023, 83, 819–824. [Google Scholar] [CrossRef] [PubMed]
  165. Di Carlo, P.; Punzi, G.; Ursini, G. Brain-Derived Neurotrophic Factor and Schizophrenia. Psychiatr. Genet. 2019, 29, 200–210. [Google Scholar] [CrossRef] [PubMed]
  166. Jauhar, S.; Johnstone, M.; McKenna, P.J. Schizophrenia. Lancet 2022, 399, 473–486. [Google Scholar] [CrossRef] [PubMed]
  167. del Pino, I.; García-Frigola, C.; Dehorter, N.; Brotons-Mas, J.R.; Alvarez-Salvado, E.; Martínez de Lagrán, M.; Ciceri, G.; Gabaldón, M.V.; Moratal, D.; Dierssen, M.; et al. Erbb4 Deletion from Fast-Spiking Interneurons Causes Schizophrenia-like Phenotypes. Neuron 2013, 79, 1152–1168. [Google Scholar] [CrossRef] [PubMed]
  168. Yang, J.-M.; Shen, C.-J.; Chen, X.-J.; Kong, Y.; Liu, Y.-S.; Li, X.-W.; Chen, Z.; Gao, T.-M.; Li, X.-M. Erbb4 Deficits in Chandelier Cells of the Medial Prefrontal Cortex Confer Cognitive Dysfunctions: Implications for Schizophrenia. Cereb. Cortex 2019, 29, 4334–4346. [Google Scholar] [CrossRef] [PubMed]
  169. Abi-Dargham, A.; Mawlawi, O.; Lombardo, I.; Gil, R.; Martinez, D.; Huang, Y.; Hwang, D.-R.; Keilp, J.; Kochan, L.; Van Heertum, R.; et al. Prefrontal Dopamine D 1 Receptors and Working Memory in Schizophrenia. J. Neurosci. 2002, 22, 3708–3719. [Google Scholar] [CrossRef]
  170. Xing, B.; Guo, J.; Meng, X.; Wei, S.; Li, S. The Dopamine D1 but Not D3 Receptor Plays a Fundamental Role in Spatial Working Memory and BDNF Expression in Prefrontal Cortex of Mice. Behav. Brain Res. 2012, 235, 36–41. [Google Scholar] [CrossRef] [PubMed]
  171. Leriche, L.; Bezard, E.; Gross, C.; Guillin, O.; Foll, B.; Diaz, J.; Sokoloff, P. The Dopamine D3 Receptor: A Therapeutic Target for the Treatment of Neuropsychiatric Disorders. CNS Neurol. Disord.-Drug Targets 2006, 5, 25–43. [Google Scholar] [CrossRef]
  172. Guillin, O.; Diaz, J.; Carroll, P.; Griffon, N.; Schwartz, J.-C.; Sokoloff, P. BDNF Controls Dopamine D3 Receptor Expression and Triggers Behavioural Sensitization. Nature 2001, 411, 86–89. [Google Scholar] [CrossRef]
  173. Harb, M.; Jagusch, J.; Durairaja, A.; Endres, T.; Leßmann, V.; Fendt, M. BDNF Haploinsufficiency Induces Behavioral Endophenotypes of Schizophrenia in Male Mice That Are Rescued by Enriched Environment. Transl. Psychiatry 2021, 11, 233. [Google Scholar] [CrossRef] [PubMed]
  174. Nieto, R.R.; Carrasco, A.; Corral, S.; Castillo, R.; Gaspar, P.A.; Bustamante, M.L.; Silva, H. BDNF as a Biomarker of Cognition in Schizophrenia/Psychosis: An Updated Review. Front. Psychiatry 2021, 12, 662407. [Google Scholar] [CrossRef] [PubMed]
  175. Fiore, M.; Angelucci, F.; Aloe, L.; Iannitelli, A.; Korf, J. Nerve Growth Factor and Brain-Derived Neurotrophic Factor in Schizophrenia and Depression: Findings in Humans, and Animal Models. Curr. Neuropharmacol. 2003, 1, 109–123. [Google Scholar] [CrossRef]
  176. Angelucci, F.; Brenè, S.; Mathé, A.A. BDNF in Schizophrenia, Depression and Corresponding Animal Models. Mol. Psychiatry 2005, 10, 345–352. [Google Scholar] [CrossRef] [PubMed]
  177. Want, A.; Morgan, J.E.; Barde, Y.-A. Brain-Derived Neurotrophic Factor Measurements in Mouse Serum and Plasma Using a Sensitive and Specific Enzyme-Linked Immunosorbent Assay. Sci. Rep. 2023, 13, 7740. [Google Scholar] [CrossRef] [PubMed]
  178. Want, A.; Nan, X.; Kokkali, E.; Barde, Y.-A.; Morgan, J.E. Brain-Derived Neurotrophic Factor Released from Blood Platelets Prevents Dendritic Atrophy of Lesioned Adult Central Nervous System Neurons. Brain Commun. 2023, 5, fcad046. [Google Scholar] [CrossRef]
  179. Klein, A.B.; Williamson, R.; Santini, M.A.; Clemmensen, C.; Ettrup, A.; Rios, M.; Knudsen, G.M.; Aznar, S. Blood BDNF Concentrations Reflect Brain-Tissue BDNF Levels across Species. Int. J. Neuropsychopharmacol. 2011, 14, 347–353. [Google Scholar] [CrossRef]
Figure 3. Trk neurotrophin receptors. (A): Structure and isoforms of the Trk receptors. Trk receptors consist of cysteine clusters (CC-1 and CC-2), leucine-rich motifs (LRMs), immunoglobulin-like C2-type motifs (Ig-C2), a transmembrane domain (TM), a juxtamembrane domain (JX), and a catalytic tyrosine-kinase domain (Tyr-K). TrkA isoforms bind NGF and differ on the six amino acid residues (VSFSPV) present only in the neuronal-specific TrkA receptor. TrkB isoforms bind BDNF and NT-4/5 and include the catalytic tyrosine kinase full-length isoforms and the non-catalytic truncated ones (T1 and T2). TrkC has three catalytic tyrosine kinase isoforms (K14, K25, and K39) and four non-catalytic truncated isoforms (TrkCTK-158, TrkCTK-143, TrkCTK-113, and TrkCTK-108) that bind NT-3. Adapted from reference [34]. (B): Schematic representation of Trk receptor catalytic isoform expression along rat development. Data acquired from northern blot assays from reference [41].
Figure 3. Trk neurotrophin receptors. (A): Structure and isoforms of the Trk receptors. Trk receptors consist of cysteine clusters (CC-1 and CC-2), leucine-rich motifs (LRMs), immunoglobulin-like C2-type motifs (Ig-C2), a transmembrane domain (TM), a juxtamembrane domain (JX), and a catalytic tyrosine-kinase domain (Tyr-K). TrkA isoforms bind NGF and differ on the six amino acid residues (VSFSPV) present only in the neuronal-specific TrkA receptor. TrkB isoforms bind BDNF and NT-4/5 and include the catalytic tyrosine kinase full-length isoforms and the non-catalytic truncated ones (T1 and T2). TrkC has three catalytic tyrosine kinase isoforms (K14, K25, and K39) and four non-catalytic truncated isoforms (TrkCTK-158, TrkCTK-143, TrkCTK-113, and TrkCTK-108) that bind NT-3. Adapted from reference [34]. (B): Schematic representation of Trk receptor catalytic isoform expression along rat development. Data acquired from northern blot assays from reference [41].
Ijms 25 08312 g003
Figure 4. p75NTR structure and interaction with other receptors. p75NTR consists of an extracellular domain (p75ECD), consisting of four cysteine clusters (CCs), a transmembrane domain (TM), and an intracellular domain (p75ICD), formed by a chopper domain (Ch-D) and a death domain (D-D). p75NTR can form homodimers or heterodimers (with Trk and sortilin) to exert different functions. Also, p75NTR can be cleaved by the α- and γ-secretases to separate its extracellular and intracellular domains. p75ICD can then bind to TrkA receptor dimers to modulate their affinity for NGF, forming a tetramer complex. Adapted from references [39,40].
Figure 4. p75NTR structure and interaction with other receptors. p75NTR consists of an extracellular domain (p75ECD), consisting of four cysteine clusters (CCs), a transmembrane domain (TM), and an intracellular domain (p75ICD), formed by a chopper domain (Ch-D) and a death domain (D-D). p75NTR can form homodimers or heterodimers (with Trk and sortilin) to exert different functions. Also, p75NTR can be cleaved by the α- and γ-secretases to separate its extracellular and intracellular domains. p75ICD can then bind to TrkA receptor dimers to modulate their affinity for NGF, forming a tetramer complex. Adapted from references [39,40].
Ijms 25 08312 g004
Figure 5. TrkB-FL (A), TrkB-T1 (B), and p75NTR (C) mRNA distribution and relative levels in murine (left) and human (right) brains during development, according to RNAseq data from reference [17]. Dimmer colors mark higher mRNA levels, relativized to the maximum expressing point for each graph. E: mouse embryonic stage; P: mouse postnatal day; PCW: human postcoital week. N/D: data not available.
Figure 5. TrkB-FL (A), TrkB-T1 (B), and p75NTR (C) mRNA distribution and relative levels in murine (left) and human (right) brains during development, according to RNAseq data from reference [17]. Dimmer colors mark higher mRNA levels, relativized to the maximum expressing point for each graph. E: mouse embryonic stage; P: mouse postnatal day; PCW: human postcoital week. N/D: data not available.
Ijms 25 08312 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hernández-del Caño, C.; Varela-Andrés, N.; Cebrián-León, A.; Deogracias, R. Neurotrophins and Their Receptors: BDNF’s Role in GABAergic Neurodevelopment and Disease. Int. J. Mol. Sci. 2024, 25, 8312. https://doi.org/10.3390/ijms25158312

AMA Style

Hernández-del Caño C, Varela-Andrés N, Cebrián-León A, Deogracias R. Neurotrophins and Their Receptors: BDNF’s Role in GABAergic Neurodevelopment and Disease. International Journal of Molecular Sciences. 2024; 25(15):8312. https://doi.org/10.3390/ijms25158312

Chicago/Turabian Style

Hernández-del Caño, Carlos, Natalia Varela-Andrés, Alejandro Cebrián-León, and Rubén Deogracias. 2024. "Neurotrophins and Their Receptors: BDNF’s Role in GABAergic Neurodevelopment and Disease" International Journal of Molecular Sciences 25, no. 15: 8312. https://doi.org/10.3390/ijms25158312

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