Next Article in Journal
Hsp70—A Universal Biomarker for Predicting Therapeutic Failure in Human Female Cancers and a Target for CTC Isolation in Advanced Cancers
Next Article in Special Issue
A Truncated Receptor TrkB Isoform (TrkB.T1) in Mechanisms of Genetically Determined Depressive-like Behavior of Mice
Previous Article in Journal
Alveolar Ridge Augmentation with a Novel Combination of 3D-Printed Scaffolds and Adipose-Derived Mesenchymal Stem Cells—A Pilot Study in Pigs
Previous Article in Special Issue
BDNF rs962369 Is Associated with Major Depressive Disorder
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Brain-Derived Neurotrophic Factor Dysregulation as an Essential Pathological Feature in Huntington’s Disease: Mechanisms and Potential Therapeutics

1
Department of Neurological Surgery, Washington University School of Medicine in St. Louis, St. Louis, MO 63110, USA
2
Department of Neurology, Washington University School of Medicine in St. Louis, St. Louis, MO 63110, USA
3
Department of Genetics, Washington University School of Medicine in St. Louis, St. Louis, MO 63110, USA
4
Hope Center for Neurological Disorders, Washington University School of Medicine in St. Louis, St. Louis, MO 63110, USA
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(8), 2275; https://doi.org/10.3390/biomedicines11082275
Submission received: 30 June 2023 / Revised: 7 August 2023 / Accepted: 9 August 2023 / Published: 16 August 2023
(This article belongs to the Special Issue BDNF in Brain Disorders: From Pathogenesis to Treatment)

Abstract

:
Brain-derived neurotrophic factor (BDNF) is a major neurotrophin whose loss or interruption is well established to have numerous intersections with the pathogenesis of progressive neurological disorders. There is perhaps no greater example of disease pathogenesis resulting from the dysregulation of BDNF signaling than Huntington’s disease (HD)—an inherited neurodegenerative disorder characterized by motor, psychiatric, and cognitive impairments associated with basal ganglia dysfunction and the ultimate death of striatal projection neurons. Investigation of the collection of mechanisms leading to BDNF loss in HD highlights this neurotrophin’s importance to neuronal viability and calls attention to opportunities for therapeutic interventions. Using electronic database searches of existing and forthcoming research, we constructed a literature review with the overarching goal of exploring the diverse set of molecular events that trigger BDNF dysregulation within HD. We highlighted research that investigated these major mechanisms in preclinical models of HD and connected these studies to those evaluating similar endpoints in human HD subjects. We also included a special focus on the growing body of literature detailing key transcriptomic and epigenetic alterations that affect BDNF abundance in HD. Finally, we offer critical evaluation of proposed neurotrophin-directed therapies and assessed clinical trials seeking to correct BDNF expression in HD individuals.

1. Introduction

Huntington’s disease (HD) is a universally fatal, autosomal-dominant neurodegenerative disease that affects nearly 40,000 individuals in the United States. This “hereditary chorea,” first fully described and reported by George Huntington in 1872, typically affects individuals in middle age and results in their death within 20 years [1,2]. Despite its debilitating nature in affected individuals, significant breakthroughs in the understanding of HD pathogenesis were not made until the mid-1990s, when the identification of a toxic gain-of-function mutation was first reported [3]. Accordingly, the clinical description of HD has advanced significantly from its initial and primary classification as a movement disorder to a complex neurological condition that affects diverse neuroanatomical substrates and most often includes cognitive and psychiatric symptoms.
The neuronal atrophy in HD is the consequence of a CAG trinucleotide repeat expansion occurring in exon 1 of the huntingtin (HTT) gene on chromosome 4, which encodes a mutant protein (mHTT) with an N-terminal polyglutamine repeat expansion [3]. In HD individuals, an expansion greater than 39 CAG repeats results in 100% penetrance of the HD phenotype, whereas a CAG repeat length between 36 and 39 may result in incomplete penetrance or late-onset of the HD phenotype [4,5]. The N-terminal repeat expansion in mHTT confers a toxic gain-of-function, as this expanded polyglutamine region imparts several new properties to the protein. The most well-known property of mHTT is the ability to form insoluble inclusions with itself and other proteins, but mHTT also disrupts numerous neuron-critical functions in manners independent of aggregate formation. Accordingly, the length of the CAG repeat in HTT generally correlates inversely with age-of-onset of HD motor symptoms [6,7,8]. The instability of an expanded trinucleotide repeat, a prominent characteristic seen in HD and other triplet repeat disorders, is associated with disease severity due to expansions in both germline and somatic cells. While the exact mechanisms of repeat instability are not completely understood, various hypotheses have been proposed, involving DNA replication, repair, and gene transcription [9]. Consequently, observations of differential repeat length expansions in the most severely affected neuroanatomical structures may help to explain the regional vulnerability to HD pathogenesis [10,11,12,13].
The profound movement dysfunction characteristic of HD is at first the result of the selective loss of subpopulations of medium spiny neurons (MSNs) of the caudate and putamen, two subregions of the basal ganglia that have a shared canonical role in the control of voluntary movements and suppression of involuntary movements. It is near-universally accepted that the initial loss of D2 dopamine receptor-expressing MSNs of the indirect pathway of movement generates the distinctive chorea in the early stages of manifest HD [14,15]. Emerging research also indicates that other factors may affect MSN vulnerability to degeneration in early HD. Recent data-driven approaches indicate that MSN vulnerability may hinge on the position within the striosome–matrix axis [16], the absence or presence of mitochondrial RNA release [17], and the occurrence of somatic repeat expansions [18]. The precise mechanisms underlying increased vulnerability associated with these factors have yet to be elucidated. Whichever factors confer the initial susceptibility to neurodegeneration in HD, the neuron loss that was at first region-specific, eventually progress to broader atrophy of the neostriatum and a majority of telencephalic structures. In fact, the Vonsattel neuropathological staging criteria recognize pathological changes to occur in numerous neuroanatomical regions at later stages of disease progression, including the broader cortical mantle, cerebellum, thalamus, and substantia nigra [19,20,21]. Loss of white matter integrity, believed to be both a cause and result of damage of gray matter neurons, is another pathological feature of HD and can be detected through multiple imaging modalities in HD individuals and small animal models [22,23,24,25].
In addition to the movement disorder manifest in HD individuals, it is also well-established that cognitive impairments and affective symptoms are highly co-morbid with HD and often occur prior to the onset of overt HD chorea [26]. This observation is not unexpected, given the role for basal ganglia signaling in higher cognitive function and emotional regulation [27,28]. At later stages of disease progression, more distant neuroanatomical structures exhibit neuronal loss, and the cognitive domains regulated by these neural substrates become altered. For example, volumetric magnetic resonance imaging reports involving HD populations and rodent models consistently detect the loss of gray matter in the hippocampus and other limbic structures [29,30,31]. These atrophic changes are frequently correlated with poorer performance on spatial memory tasks in HD individuals and on hippocampal-dependent behavioral assays in rodent models [32,33,34]. The psychiatric burden in the HD population is substantially higher than for most neurodegenerative disease patient populations, with some observational reports and meta-analyses suggesting lifetime prevalence of psychiatric disorders to exist between 33 and 76% [35,36,37]. These psychiatric disturbances often occur well before the onset of the overt dysfunction of motor control and negatively contribute to quality of life in HD individuals [38]. Exploration of the shared features causal to the motor, cognitive, and psychiatric symptoms that manifest in HD is likely to illuminate common mechanisms amenable to intervention in this patient population.
The examination of molecular features associated with mHTT expression has generated significant insight into specific cytological disruptions present at multiple stages of HD and in corresponding HD models. These include impairments of the regulation of gene expression [39], translation [40], autophagy [41,42], mitochondrial function [43], intraneuronal transport [44], and synaptic function [45]. These impairments possess independent damaging effects on neuronal health and function, but they can also promote detrimental additive effects on a neuron’s ability to withstand the mounting burden of neurotoxic events in HD. As no disease-modifying therapeutic options currently exist for the treatment of HD, the identification and dedicated study of neuron-protective factors are paramount. There is perhaps no greater example of a set of broadly protective factors in the context of neurodegenerative environments than neurotrophins and neurotrophin receptors. In this literature review, we discuss neurotrophins, the compromise of their function in HD, and how their expression or regulation might be manipulated to delay disease progression in HD, with a special focus on the brain-derived neurotrophic factor (BDNF).

2. BDNF and TrkB-BDNF Signaling: Critical Intersections with Common Neurodegenerative Processes

BDNF is a member of the neurotrophin family, a small but critically important series of diffusible neuropeptides with countless roles in neuronal development, survival, and function, including synaptic plasticity. Neurotrophins engage major neuronal growth and anti-apoptotic signaling cascades through action at either tropomyosin-receptor kinase (Trk) or Trk-p75NTR heteromeric receptor complexes. For this reason, exposure to neurotrophins, whether constitutive or activity-associated, is strongly supportive of neuron structure and function and can often determine a neuron’s capacity to withstand exposure to neurotoxic stimuli. BDNF and BDNF-responsive neurotrophin receptors are reviewed extensively in this special issue and in recent literature reviews [46,47,48]. Nonetheless, understanding of common features of BDNF signaling outcomes within neurodegenerative processes will serve to illuminate the critical role of BDNF in HD pathogenesis and inform potential therapeutic strategies for the treatment of HD and related conditions.
Landmark studies in multiple neurodegenerative fields have repeatedly demonstrated shared impairments of neurotrophin homeostasis in human subjects and rodent models. These impairments most often include the regional downregulation of BDNF expression in affected neural structures [49,50] or in cerebrospinal fluid (CSF) vs. controls [51,52,53,54]. This downregulation is often associated with the dysregulation of transcription and translation of the BDNF gene, but it may also exist as a result of disruption of the proteolysis of its precursor, proBDNF, or through the altered metabolism of its mature isoform [54,55]. To compound the loss of BDNF expression, the trafficking and subsequent release of BDNF is frequently compromised as neural structures degrade, leading to decreased constitutive or activity-dependent release of neurotrophic factors [44,56]. In this way, early loss of neurotrophic support in neurodegenerative diseases, especially BDNF signaling, often feeds into a vicious cycle of molecular events: reduced BDNF signaling impedes the neuronal ability to maintain functional transport systems and synaptic connections, which in turn decreases both the release of BDNF and the response to its presence through its receptors. Nerve growth factor (NGF) and neurotrophins 3/4 (NT3, NT4) also contribute to neuronal survivability, but their action in the mature (i.e., post-development) CNS appears somewhat restricted to special neuroanatomical regions with high expression of their cognate receptors. For example, the loss of supportive NGF signaling in the basal forebrain remains a favored hypothesis for the early and severe regional degeneration of cholinergic cells in Alzheimer’s disease, but NGF alone may be insufficient to explain the degeneration observed in forebrain-innervated regions [57]. Nonetheless, the broader expression of BDNF and greater CNS abundance relative to other neurotrophin family members suggest that the investigation of BDNF dysfunction in neurodegenerative diseases is likely to offer mechanistic insight and fruitful avenues for therapeutic advances.
The Trk family of neurotrophin receptors has likewise attracted considerable research attention for its role as the major signal transducer system for mature neurotrophins. While the affinity between neurotrophins and neurotrophin receptors is somewhat promiscuous, the strongest bindings (i.e., lowest KD) are largely acknowledged to occur between NGF:TrkA, BDNF and NT4:TrkB, and NT3:TrkC complexes. The activation of Trk family receptors is anti-apoptotic in nearly all contexts through the activation of major cell signaling pathways in neurons, including, but not limited to, PI3K/AKT and Ras/MAPK/ERK [58]. These Trk-associated signaling pathways, when activated, also serve to modulate synaptic plasticity [59], neurite regeneration [60], and neurotransmitter release [61]. A fourth member of the neurotrophin receptor family, the p75 neurotrophin receptor (p75NTR), forms Trk-p75NTR heteromeric complexes with an exceptionally high affinity for neurotrophins [62]. In an opposite manner, p75NTR also forms a co-receptor complex with sortilin and drives pro-atrophic signaling through the high-affinity binding of proneurotrophins [63]. For this reason, the investigation of changes in Trk abundance is rarely considered in isolation but must be evaluated in light of corresponding changes in p75NTR and pro- and mature neurotrophin expression. The marked and broad downregulation of Trk expression has nevertheless been described in numerous research reports involving major neurodegenerative diseases [64,65,66]. These studies, both in humans and preclinical models, suggest that TrkB loss is often an early event in neurodegenerative disease progression and accompanies the simultaneous downregulation of BDNF [67,68]. Additional reports indicate that TrkB activity may also be inhibited in these progressive disease states via a mechanism involving the direct binding of neurotoxic peptides to internal TrkB epitopes (e.g., alpha-synuclein) or the differential endogenous expression of Trk adaptor proteins [50,69]. Curiously, alternative splicing of TrkB into catalytically inactive “truncated” isoforms (e.g., TrkB.T1) may also act to throttle the broader Trk response to the existing BDNF pool through dominant-negative inhibition. The exploration of the differential expression of full vs. truncated isoforms in models of neurodegenerative diseases represents an exciting research direction, especially in light of reports showing that the manipulation of the TrkB:TrkB.T1 ratio acts to modify disease onset [70,71,72].
Finally, the up-regulation of the pan-proneurotrophin receptor p75NTR is a consistent observation in the context of both acute and chronic neural injury [73]. For this reason, impaired processing from the precursor to the mature isoform of BDNF also serves as an explanation for accelerated neuronal death, even when Trk receptor expression remains stable. Mechanisms underpinning the disruption of (pro)neurotrophin processing appear to be more disease-specific than shared, but they may include the downregulation of cellular proteases [55,74] or the modification of proconvertase activity by the expression of inhibiting peptides [75]. Imbalances in the Trk/p75NTR ratio to favor increased p75NTR signaling are also described in multiple research reports, and some of these studies further indicate that this ratio may be affected at a subcellular or synaptic level [64,76]. Whichever mechanisms predominate, the neurotrophin and neurotrophin receptor signaling network exemplifies a dynamic ligand-receptor system, whose disruption has outsized effects on neuronal survival and function in the context of neurodegenerative disease.

3. The Role of BDNF in HD

Since their discovery, neurotrophins have been inextricably linked to neuron viability in the face of both age- and disease-associated pathogenic stimuli. For this reason, impairments of BDNF expression, signal transduction, transport, and metabolism remain fertile targets for investigation within progressive neurological disorders. HD in particular serves as a model neurodegenerative disease in which the normal suites of neurotrophin and neurotrophin receptor functions are compromised. This disruption is repeatedly observed in HD individuals and nearly all available preclinical models; therefore, it is likely to both serve as a critical factor modifying disease progression and simultaneously offer mechanistic insight into major cellular features manifest in HD subjects.

3.1. Loss of BDNF in HD

BDNF expression and biodistribution in Huntington’s brains are dramatically affected in the course of disease progression. Early reports examining BDNF expression in post-mortem tissue from HD individuals found a reduction of up to 82% of BDNF expression in the caudate and putamen vs. controls [77]. Landmark studies by two additional research teams indicated that decreased cortical BDNF gene transcription [78] and interruption of the expression of BDNF vesicular transport proteins [79] were both present in HD brains (discussed below). These two findings appear especially critical with respect to HD pathogenesis, as the vast majority of striatal BDNF is synthesized in cortical neurons and subsequently transported to targeted striatal regions [80,81]. TrkB expression is similarly affected in the caudate nucleus and motor cortex of post-mortem tissue from HD brains, with some striking reports indicating up to an 80% loss of TrkB immunoreactivity in HD vs. control brains [82]. In a somewhat contradictory set of findings, the major Val66Met polymorphism in the pro-domain of BDNF, a significant negative modifier of disease progression in several neurodegenerative, affective, and psychiatric disorders, appeared to have no effect on HD pathogenesis [83]. In fact, one report indicated that HD individuals who were heterozygous for this polymorphism exhibited a later age-of-onset than those homozygous for valine or methionine at residue 66, although this effect was only observed in a subset of HD subjects in their study [84].
At least two reports indicated that serum BDNF is decreased in HD patients, which appears to correlate inversely with CAG length and the duration of illness, as measured by the unified HD rating scale (UHDRS) [85,86]. At least one ongoing clinical trial evaluated BDNF or TrkB in CSF from HD individuals as a biomarker of disease progression (NCT04012411). More recent reports, however, conflict with earlier observations of decreased peripheral BDNF abundance in HD and further suggest that BDNF alone may not represent as a suitable biomarker for disease progression [87,88]. Ultrasensitive ELISAs do not identify a difference in CSF BDNF between controls and HD individuals, even in advanced-stage HD subjects [88]. This finding is somewhat puzzling in light of the strong positive correlation of measures of disease progression with other CNS-specific biomarkers, including CSF tau [89], mHTT [90], and neurofilament light chain [91,92], but not unexpected given the short BDNF half-life and multiple non-neural pools of neurotrophin expression. The interpretation of plasma or CSF BDNF may be further complicated by the likely compensatory increase in its expression by peripheral cell types. In fact, it was recently reported that a moderate increase in intra-platelet BDNF levels was detected in patients in stage II of HD [93]. Investigation of alternative peripheral pools of BDNF that mirror the decreases occurring in neural structures continues to be of importance in HD [94].

3.2. BDNF Interruption in Preclinical Models and Mechanistic Insights

A decline in BDNF expression is not only observed in HD individuals but is reproduced consistently within HD models. In fact, the expression profiling of striatal tissue obtained from two HD models, when compared to humans, indicated that BDNF depletion may represent the most critical role underlying striatal degeneration in HD individuals [95]. In agreement with this finding, rapidly progressive models of HD, such as the R6/2 transgenic mouse, showed an age-dependent reduction in BDNF expression and rescue of TrkB-associated signaling via the administration of recombinant BDNF [96]. The overexpression of BDNF and TrkB in an immortalized cell line model also restored TrkB activity and attenuates caspase-3 activation [97]. Further, striatal BDNF reduction in zQ175 mice was found to exist without corresponding decrements in proBDNF, indicating that the loss of BDNF abundance in HD may involve both transcription-dependent and -independent mechanisms [98]. Interestingly, reduced TrkB activity preceded the loss of BDNF in some of these models, suggesting that impaired TrkB-mediated signaling represents the first event in the cascade of neurotrophin dysregulation events in HD [50,99].
Curiously, loss of BDNF is even observed in neuroanatomical structures largely spared by classical HD pathology, although it is unknown if the majority of these observations are restricted to models. In both HD individuals and the N171-82Q model, BDNF expression is sharply reduced in the pontine and medullar regions of the brainstem and was suggested to contribute to accelerated heart rate. Intracerebroventricular administration of BDNF was sufficient to restore the heart rate to normal levels in these HD mice [100]. Disruption of hypothalamic function, believed to generate circadian dysfunction and neuroendocrine abnormalities in HD individuals, may also be related to the suppression of resident BDNF and/or TrkB signaling [101,102]. For example, the breakdown of circadian rest-activity cycles was observed in mouse models and compared well with robust sleep disturbances in HD individuals. The interruption of BDNF-TrkB signaling in the suprachiasmatic nucleus, a hypothalamic subregion responsible for regulating light-dark cycle entrainment, may serve as a contributing factor, especially due to the high local expression of these two proteins [103].
Modeling of HD within in vitro systems largely mirrors the decline in BDNF abundance in rodent models and HD individuals. Indeed, mouse primary neuron models that employ the lentiviral transduction of mHTT exon 1 display the downregulation of Bdnf gene expression well before decrements in measures of viability [104,105]. Untransfected primary neurons harvested from embryonic-age HD mouse models also displayed aberrations in BDNF expression, transport, or release in culture [106,107,108]. More innovative approaches, including induced pluripotent stem cells (iPSCs) and the direct reprogramming of fibroblasts obtained from HD patients, are increasingly used to model HD [109]. BDNF appeared within dysregulated gene networks in some iPSC models [110], and the HD iPSC-derived neural progenitor cells appeared to be more vulnerable to BDNF withdrawal than control iPSC-derived cells [111]. In support of these results, reprogrammed striatal neurons from HD fibroblasts displayed the downregulation of gene expression of the mRNA encoding TrkB vs. striatal neurons reprogrammed from unaffected individuals [112]. Another recent publication employing HD dermal fibroblasts directly reprogrammed to striatal-lineage neurons failed to show differential BDNF or NTRK2 (TrkB) gene expression between HD and control groups, but HD neurons displayed neurite dysmorphogenesis and electrophysiological abnormalities—two neuronal attributes associated with neurotrophin signaling [113]. Finally, limited reports evaluating BDNF endpoints within immortalized striatal neuron models also suggested that neurotrophin abnormalities also exist in these cell lines. STHdhQ111/111 striatal cells displayed both reduced expression and subcellular localization of TrkB and impaired BDNF-induced MAPK signaling vs. cells of the STHdhQ7/7 control line [114]. Ultimately, the collection of these model systems not only serves to describe the extent of BDNF loss in HD, but it aids in uncovering the set of common mechanisms underpinning BDNF disruption and illuminates the intersections amenable for therapeutic approaches.
HTT’s role as a scaffolder of molecular motor complexes and their cargoes places it in a pivotal intersection to disrupt BDNF’s functions at distal neuronal structures, as anterograde and retrograde axonal transport is critically linked with neurotrophin release and signal transduction. Microfluidic and real-time imaging approaches employing neurons obtained from HD models have been particularly illuminating in examining trafficking deficits in mHTT-expressing neurons. Indeed, a decrease in BDNF travel length and the speed of BDNF-containing vesicles was observed in primary neurons obtained from zQ175 HD model mice [106]. BDNF-specific trafficking impairments were replicated in HdhCAG140 neurons [44] and other models [107]. A correction of transport defects in models is often sufficient to reverse major pathological features [115], indicating that restoring the axonal transport capacity in HD may represent a more attractive and practical approach than the replenishment of the BDNF level more broadly. How might mHTT disrupt axonal transport? HTT is known to serve in a trafficking complex with huntingtin-associated protein 1 (HAP1) and directly facilitates complexing and assembly of dynein and dynactin subunits [116]. Interestingly, strength of the HTT-HAP1 interaction was found to be positively correlated with polyglutamine repeat length, which is likely to impact the trafficking of BDNF- or Trk-laden vesicles [117]. Retrograde transport of TrkB is also negatively affected in a polyglutamine-dependent manner—a finding which has particular implications for the fidelity of TrkB retrograde signal transduction [118]. Finally, the phosphorylation state of mHTT, particularly at serine 421, appeared to modulate the ability of mHTT to bind subunits of motor complexes and transport intracellular cargoes, including BDNF [119]. Phosphoproteomic alterations in neurons, a striking recent finding in pre-symptomatic HD mice, might therefore serve as interesting indirect mechanisms that have knock-on effects on BDNF transport [120].

3.3. Effects of BDNF Loss on Synaptic Features

There is perhaps no greater purpose associated with BDNF than its repertoire of roles at the synapse, where BDNF signaling is firmly established as a critical factor shaping synaptic development, structure, and function [121]. Loss of synaptic integrity is well-established to underpin deleterious effects on cognitive, behavioral, and motor functions [122]. Consequently, the early observations of aberrant BDNF expression in HD individuals and models quickly led to multiple lines of investigation focusing on synaptic integrity and impairments in synaptic transmission in HD.
There are likely effects of mHTT on synaptic function that may be mediated by the loss of synaptic BDNF expression or involve additional knock-on effects of altered BDNF-TrkB signaling in HD. Early evidence for a causal role for BDNF in deficits of synaptic efficacy was provided by two groups demonstrating that the overexpression or exogenous administration of BDNF in HD models partially rescues deficits in long-term potentiation (LTP), a major model of synaptic plasticity [123,124]. Xie and colleagues additionally demonstrated that decrements in multiple molecular and behavioral phenotypes in the YAC128 model were ameliorated when crossed with a transgenic line, which overexpressed BDNF in forebrain neurons [125]. Moreover, the pharmacological induction of BDNF through α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPAR) modulators was sufficient to rescue LTP deficits and activity-driven actin polymerization at the dendritic spine in a CAG140 mouse model [126]. The restoration of plasticity-associated endpoints via the exogenous application of neurotrophins is a Trk isoform-dependent process and appears to be a unique property of BDNF among neurotrophins [127]. The stimulation of adenosine-2A receptors, believed to directly activate TrkB receptors and potentiate the BDNF-mediated modulation of synaptic transmission, also appeared to restrain excitotoxicity in R6/2 rodents [128,129]. The expression of mHTT is not merely associated with reduced plastic capacity within LTP paradigms, but also affects the activity-dependent release of BDNF into the synaptic cleft, at least at corticostriatal synapses. A unique approach involving the real-time sensing of BDNF release revealed the interruption of fusion modes of BDNF-laden vesicles following the high-frequency stimulation of HdhQ175-expressing cortical projection neurons [130]. Finally, mHTT expression may also be associated with the maladaptive recovery of synaptic function after neural injury. A recent study by Spies et al. indicated that induced dsDNA breaks in the STHdhQ111/111 cell model had differential effects on gene expression associated with synaptic plasticity vs. controls, including a lack of a compensatory Bdnf and Arc upregulation [131].
There is limited but exciting evidence that the synaptic expression of neurotrophin receptors is similarly compromised in HD. Reduced expression of TrkB in HD mouse models and in HD individuals was first reported nearly two decades ago, although it was not clear whether this decrement was neuron-wide or specific to synaptic terminals, a critical site of TrkB expression [82]. Evidence of disrupted TrkB retrograde vesicular transport in striatal dendrites was shown to exist alongside reduced ERK and c-Fos activation through a unique microfluidic approach [118]. Nearly simultaneously, Brito and colleagues reported the upregulation of p75NTR at the synapse of hippocampal neurons in two mouse models of HD [132]. This upregulation of p75NTR therefore likely co-exists with decrements in synaptic TrkB expression and favors the activation of pro-atrophic/pro-apoptotic neuronal signaling cascades in humans and models [76]. When considered alongside BDNF-related abnormalities, expression and localization changes in these BDNF-responsive receptors likely exacerbate deficiencies in neurotrophin signaling associated with mHTT expression.
It is important to note that various research groups have identified roles for HTT at the synapse that are independent of BDNF, including the scaffolding of multiple cytoskeletal and synaptic vesicle proteins and the regulation of vesicle endocytosis [133,134,135]. More recent reports point to emerging roles for HTT in the modulation of vesicular protein–protein interactions, presynaptic homeostasis, and activity of terminal secretases [133,136,137]. Unsurprisingly, HTT was even shown to be required for short and long-term potentiation at Schaffer collaterals, emphasizing the importance of synaptic HTT expression [138]. Therefore, the loss of a functional HTT allele in HD likely contributes to the simultaneous exacerbation of decrements of BDNF expression and neurotrophin signaling at the synapse. In agreement with this observation, many pathological processes are reversed by the overexpression of wild-type HTT protein, indicating that loss of normal HTT function in adulthood may promote an equally damaging outcome as the expression of its mutant counterpart [139,140].

4. Mechanisms of BDNF Gene Regulation in HD

4.1. HTT-Associated Regulation of Bdnf Transcription

A central molecular feature of HD pathogenesis is the robust transcriptional dysregulation observed in affected cells, which occurs even in premanifest stages of HD [39,141]. For this reason, mechanisms favoring the altered regulation of gene expression have provided multiple ripe targets for HD investigators over the last two decades. As a direct result of the -omics era explosion of analytical techniques, details of HTT/mHTT gene regulatory networks have now grown to staggering scales. Wild-type HTT is known to interact with several transcription factors (TFs) [142], but a mechanistic explanation for transcriptomic dysregulation involving HTT loss-of-function as a scaffolder of transcription complexes is not yet supported. In fact, a study employing neural tissue from various HD models did not show a main effect of wild-type HTT gene dosage on the degree or number of transcriptional alterations [143]. Instead, mHTT likely adopts a toxic gain-of-function through the sequestration or segregation of TFs within inclusion bodies. In support of this hypothesis, there exist several reports showing that the induced expression of protein fragments with polyglutamine expansions is sufficient to sequester major gene regulatory proteins within nuclear inclusion bodies [144,145]. Alternatively, given that HTT and mHTT occupy different genomic loci, existing gene regulation machinery may be routed to disparate genomic regulatory regions by mHTT and inappropriately suppress or permit the synthesis of corresponding transcripts [146,147]. The decreased transcription of BDNF is a central feature of HD molecular pathology [78]. Consequently, gene regulation of this neurotrophin provides an ideal model for the examination of genes whose expression is interrupted in HD.
The genomic structure of the BDNF gene in both rodents and humans is extraordinarily complex [148]. This intricate genomic architecture, however, permits the expression and distribution of Bdnf transcripts by neurons in precise spatiotemporal patterns [149]. At least nine exons exist within the Bdnf gene. The Bdnf architecture includes multiple 5′ noncoding exons and one exon containing the pre-proBDNF coding sequence. Each of these 5′ non-coding exons is linked to separate promoters. Additionally, multiple untranslated regions downstream (3′-) of the Bdnf coding sequence are known to be involved in the regulation of transport of the mature transcript [150,151]. This complex exon-promoter scheme allows for a remarkable degree of control of expression of gene regulation, but it also serves as a critical vulnerability in disease states involving the altered regulation of gene expression; at least three exons (exons II, IV, and VI) are known to be downregulated in human HD [78,152,153]. BDNF gene expression in neurons is partially, but directly, regulated by wild-type HTT through its interaction with a transcriptional repressor, the repressor-element 1-silencing transcription/neuron-restrictive silencer factor (REST/NRSF). Wild-type HTT sequesters REST/NRSF in the cytoplasm, preventing it from binding to a corresponding response element or neuron-restrictive silencer element (RE1/NRSE) within promoter II of Bdnf, thereby facilitating Bdnf transcription [152,154]. Mutant HTT, on the other hand, fails to sequester NRSF/REST, accelerating its translocation to the nucleus where it binds to the BDNF promoter and represses its transcription (Figure 1A) [155]. In support of this mechanism, the expression of a REST decoy oligonucleotide in a cellular model of HD was sufficient to alter the REST occupancy of its response element in the Bdnf gene and rescue Bdnf expression [156]. Notably, mHTT interacts with the cAMP response element-binding protein (CREB) binding protein (CBP). CBP is a transcriptional coactivator with intrinsic histone acetyltransferase (HAT) activity that binds to CREB, a transcription factor whose phosphorylation is a direct downstream outcome of TrkB activation [157]. CBP was demonstrated to be sequestered to mHTT intranuclear inclusions in HD mice [158,159]. In this way, mHTT could serve to both suppress Bdnf gene expression and diminish the neuronal response to existing BDNF-TrkB signaling through CBP/CREB. Finally, it is important to recognize that mHTT is able to bind directly to DNA and occupy genomic regulatory regions in vivo in a polyglutamine-dependent manner [146,147], providing a non-specific but key mechanism by which BDNF transcription could be negatively affected.
In addition to mHTT-associated alterations in the rate of Bdnf transcription, altered regulation of the transport of the Bdnf transcript itself may be compromised in HD. Discrepancies in mRNA transport are likely to have outsized effects in neurons, especially considering that the local and rapid translation of the Bdnf transcript is of high importance for the structure and function of dendritic spines and other distant structures. Intraneuronal Bdnf mRNA transport is orchestrated through regulatory regions within at least two distinct promoters and the 3′UTR regions of its gene body and could therefore be affected by broader patterns of regulatory disruption discussed above [151,161]. Bdnf mRNA targeting to either somatic or dendritic compartments is determined through polyadenylation states within its 3′ regulatory region [151]. Consequently, aberrant polyadenylation may serve as a contributing factor for the disruption of BDNF transcript targeting in HD, especially given recent reports of shared polyadenylation abnormalities across major neurodegenerative disorders [162]. Although no significant evidence exists to support the specific disruption of trafficking of the Bdnf transcript in HD, we note that others have shown that BDNF and its transcripts are co-segregated in neurons [161]. Therefore, reports of altered BDNF protein distribution and transportation across degrading trafficking networks within HD neurons may be equally applicable to its transcript [79,108,118].

4.2. Epigenetic Disruption of BDNF Expression

There exist many means by which epigenetic mechanisms can affect gene expression. Laying “above” (epi-) the genetic sequence encoded within the DNA itself, this series of transcriptional control mechanisms serves to affect gene transcription in both heritable and non-heritable manners. Histone post-translational modification (PTM) and DNA methylation, two forms of epigenetic regulation, have both been connected to altered expression of genes in HD. Therefore, the robust and profound disruption of gene expression in HD may arise from the failure of epigenetic control in affected neuronal populations. A greater understanding of epigenetic regulation and its failure in HD illuminates several potential areas for research and therapy.
Histone PTMs are the set of covalent alterations that occur at certain residues on the N-terminal tails of histone octamers, the core structural components of nucleosomes. These reversible modifications, which may include acetylation, ubiquitination, methylation, phosphorylation, and SUMOylation, serve to define gene expression [163]. For example, histone PTMs represent readable markers for epigenetic effector complexes and subsequently alter local chromatin states to define gene expression. Trimethylation (me3) at lysine (K) 27 of histone H3 (H3K27me3), a well-known repressive histone mark, serves to compact nearby chromatin through action of the polycomb-repressive complexes [164]. Alternatively, the ionic properties of these PTMs are also believed to broadly affect the affinity of the histone-DNA interaction and may affect gene transcription via steric accessibility and electrostatic shielding effects. In general, the repressing or activating properties of histone marks generally depend on the amino acid residue, which is modified, the type of modification, and their genomic locations, including their proximity to transcriptional start sites.
How might histone modifications be altered in HD to contribute to gene expression dysregulation, including that of BDNF? Research examining histone PTM ratios in humans and models has consistently reported the differential expression of both repressive [165,166] and activating [167,168,169,170] histone marks. The altered expression or activity of the enzymes that catalyze these histone PTMs may thereby affect PTM abundance or destabilize the ratio of repressive vs. activating marks. Inhibition of the activity of histone-modifying enzymes or the removal of the modification(s) that they deposit, however, does not consistently appear to rescue the loss of BDNF transcription in HD, even when the expression of other neuron-critical genes is restored. A study has reported the enrichment of an active histone PTM (H3K4me3) at BDNF regulatory regions in mouse and human brains, which were downregulated in the HD context, and a corresponding rescue in gene expression was observed when the demethylase catalyzing this mark was inhibited in a Drosophila model [169]. The acetylation of H3K27 is associated with Bdnf promoter occupancy and regulates Bdnf transcription [171]; however, only limited studies involving this and other marks have been conducted in HD models and humans, and these studies did not specifically evaluate BDNF expression [167,168]. The inhibition of histone deacetylases (HDACs) might also have indirect benefits on Bdnf expression by rescuing transport deficits through non-epigenetic mechanisms that target non-histone substrates. For example, HDAC6 inhibition, by increasing the levels of α-tubulin acetylation, has been demonstrated to rescue axonal transport deficits in other neurodegenerative diseases and was recently evaluated in two HD preclinical models [172,173]. The lack of additional direct evidence in this direction may indicate that the presence or absence of a particular histone PTM has only modest effects on BDNF expression in HD models. Indeed, both focused molecular and larger-omics approaches evaluating transcriptomic endpoints under the non-selective inhibition of HDACs failed to detect (or did not report) changes for Bdnf expression [174,175,176]. More likely, changes in histone PTM distribution and abundance alone do not affect Bdnf transcription, but act via the engagement of other gene expression mechanisms, including via coupling with alternative epigenetic regulatory processes. In fact, gene expression analyses of progeny from HDAC inhibitor-treated male HD mice showed improved HD phenotypes vs. vehicle controls through the mechanism involving altered DNA methylation patterning [177]. The inhibition of the families of enzymes that catalyze histone PTMs may ultimately prove to harbor beneficial effects via crosstalk between epigenetic regulatory modes. In this way, histone PTMs and DNA methylation cooperate to regulate gene expression to establish stable or dynamic chromatin states.
Methylation of the fifth position of the cytosine ring (5-methylcytosine or 5mC) is a well-characterized epigenetic mark that is associated with gene repression when present at promoters and regulatory regions. Approximately 75% of the CpGs (cytosine-guanine dinucleotides) in neurons possess the 5mC mark [178]. “CpG islands”, the unmethylated genomic regions present around regulatory regions of active genes, typically do not exhibit this mark. Thus, methylation states of promoters affect the binding of regulatory complexes to define the rate of gene transcription. As a consequence, aberrant methylation patterning can affect gene expression in pathological states. DNA methyltransferases (DNMTs), the set of enzymes that catalyze the addition of this methyl group, are present in neurons and can be active in maintenance or de novo methylation. Therefore, changes in the global methylation landscape and the up- or down-regulation of methyltransferases may both serve as mechanistic descriptions for transcriptomic aberrations in HD. In support of this hypothesis, early studies in this direction employing reduced representation bisulfite sequencing and methylated DNA immunoprecipitation (MeDIP)-seq indicated differential methylation at promoter regions in striatal cells from an HD model vs. wild-type controls [179]. Profound cytosine methylation changes also appear to be associated with mHTT expression in rodent and large animal models and HD individuals [180]. The complex genomic structure of Bdnf includes multiple promoters, and at least promoter IV is differentially methylated in HD model cortical neurons vs. controls, providing a putative but direct explanation for dysregulated Bdnf transcription (Figure 1B) [104,105]. In this model, the critical activity-associated promoter IV is directly methylated by neuronal DNMTs, while promoter VI remains unmethylated, indicating that promoter VI is not a direct target of DNA methylation [104]. Indeed, the pharmacological inhibition of DNMTs restored methylation at promoter IV of Bdnf and its transcription in vitro as well as rescued transcription of critical HD-associated genes in R6/2 mice [104]. This effect was also reproduced with the RNAi-mediated knockdown of DNMTs in HD models, suggesting a direct role for DNMTs in the hypermethylation of downregulated genes in HD. Altered BDNF promoter IV methylation in HD is not restricted to neurons in affected neuroanatomical structures, but it is also observed in non-neural cell types and may contribute to alter peripheral pools of BDNF to exacerbate the central loss of this neurotrophin [94,181]. Several intriguing reports assessing epigenetic endpoints in serum or post-mortem samples from humans suffering from other neurological and psychiatric disorders also observed the differential methylation of BDNF promoters [182,183,184]. The precise mechanisms by which DNMTs are inappropriately routed to specific promoters in HD are still unclear and require further investigation. It is important to consider that readings of the set of cytosine modifications in HD may also exist as contributing factors in epigenetic disruption. Intriguingly, HTT was found to possess a direct binding property to Methyl-CpG Binding Protein 2 (MeCP2), a critical scaffolding protein that serves as a reader of 5mC [160]. This property was shown to strengthen in a polyglutamine repeat-dependent manner, which altered the binding of MeCP2 to promoter IV of Bdnf, and resulted in its downregulation in the STHdh111/111 cell model of HD (Figure 1B).
5mC is not removed directly in post-mitotic cells but via a series of ten-eleven translocation (TET1)-dependent enzymatic conversions which include 5-hydroxymethylcytosine (5hmC), 5-formylcytosine, and 5-carboxylcytosine [185]. These alternative methylation states also appear to have gene-repressive properties at promoter regions; however, they are recognized by different reader proteins, and their roles in pathological states remain unclear. Landmark studies nearly 15 years ago established that 5hmC expression is highest in the central nervous system and comprises between 0.3–0.7% of all nucleotides in neural tissue [186,187]. Intriguingly, research teams assessing genome-wide 5hmC maps in other pathological states have reported alterations in 5hmC abundance within critical gene bodies, including that of Bdnf [188,189]. A study evaluating post-mortem HD striatal tissue observed altered 5hmC abundance in the adenosine A2A receptor (A2AR) gene body. This outcome was recapitulated in the R6/1 and R6/2 models, which correlated with decreased striatal expression of A2AR in these mice [190]. This finding is particularly intriguing, given that A2AR is a receptor whose activity is known to maintain BDNF levels in the CNS and can stimulate TrkB expression [191]. In fact, it appears that the genome-wide loss of 5hmC in the mouse brain was a key feature of at least one HD model (YAC128) at multiple ages [192]. Prenatal 5hmC loss may impart legacy effects in HD as a result of 5hmC’s proposed role in the neurodevelopmental control of gene expression [193]. In support of this observation, Barnat et al. recently showed that mHTT alters human neurodevelopment at even embryonic stages [194]. The extent to which 5hmC abundance is altered within the BDNF gene throughout developmental timeframes remains an unexplored research question that may have exciting implications, given the wide array of BDNF neurodevelopmental functions.

4.3. Other Putative Regulatory Modes for BDNF Expression

Still, other intriguing explanations have been proposed for the decrements in BDNF gene expression in HD. Splicing of pre-mRNAs into mature transcripts has only recently been shown to be broadly dysregulated in HD, an observation which could have a significant impact on Bdnf and genes with similarly complex genomic structure. Research reports detailing exon skipping, cryptic exon exposure, and intron retention are not infrequently reported in HD brains and models [195,196,197]. Intriguingly, the mHTT gene itself (but not wild-type HTT) is mis-spliced in HD individuals and knock-in models, which results in the synthesis of a transcript encoding only the pathogenic exon 1 fragment [198,199]. This finding aligns with the significant disruption of splicing factor expression observed in HD and the observations of profound splicing errors in other neurodegenerative diseases. To date, no splicing aberrations have been reported with respect to the Bdnf transcript in HD models. A growing body of research, however, has illuminated mHTT inclusion sequestration and mis-localization of the activity of the well-known splicing factor transactive response DNA-binding protein of 43kDa (TDP-43) in HD [200,201]. TDP43 was recently linked with impaired activity-dependent BDNF secretion, and it’s interruption in HD may offer a causal explanation for altered BDNF regulation through aberrant splicing [202].
MicroRNAs (miRNAs), owing to their critical post-transcriptional regulatory role, have come under particular scrutiny for a putative role in BDNF silencing [203,204]. Some reports indicate that up to 85 distinct deregulated miRNAs may exist in human HD brains [205,206]. A unique hybrid computational approach by Muller further indicated that BDNF lies at a critical hub of the miRNA-mRNA regulatory network formed by these miRNAs and, intriguingly, that two of these miRNAs (miR-10b-5p and miR-30a-5p) have putative targets in the 3′UTR of the BDNF transcript [207]. The upregulation of miR-10b-5p and miR-30a-5p, also observed in HD brains, was further predicted to strongly repress BDNF expression through an in silico approach but has not yet been demonstrated in living systems [207]. Other miRNAs may have indirect, but beneficial, effects on Bdnf expression in HD models through the restoration of other modes of gene regulation discussed above [203]. Antisense inhibition of BDNF transcription can also serve to restrain BDNF transcription in vitro and in vivo through the expression of long non-coding natural antisense transcripts (NATs). Knockout of the NAT for Bdnf can enhance cognitive function through the disinhibition of the antisense-mediated decay of the Bdnf transcript, but the manipulation of NAT expression has not yet been attempted in the context of HD [208,209].

5. BDNF-Focused Therapeutic Approaches for HD

To date, no single therapeutic strategy has shown to be effective in the reduction or delay of classical signs of HD progression in humans. The existing tetrabenazine-based FDA-approved pharmacotherapies for use in HD largely target the reduction of choreic movements in HD through the inhibition of the vesicular monoamine transporter 2 (VMAT2), but they are insufficient to mitigate HD pathophysiology [210]. The identification of additional druggable targets, including those directly causal to disease pathogenesis, is therefore desperately needed for the treatment of this devastating disease.

5.1. BDNF Supplementation

For some time, the utilization of BDNF supplementation or direct treatment has been considered as a therapeutic approach for addressing neurodegenerative disease. Accordingly, many of these approaches have been evaluated in HD models. Limited preclinical research indicates that chronic systemic BDNF delivery through indwelling osmotic pumps increased the synthesis of BDNF in the CNS and diminished some measures of disease progression in R6/2 mice [96,211]. A study by da Fonseca and colleagues demonstrated that intranasal administration of BDNF alleviated anhedonic and depressive-like behaviors at early disease stages in the YAC128 model and further stimulated hippocampal neurogenesis [212]. Unfortunately, the systemic administration of BDNF is unlikely to represent a viable neurotrophin-focused strategy in humans, as factors related to blood–brain barrier permeability, short half-life, and off-target effects in the peripheral nervous system and other structures preclude its effectiveness in vivo [213,214]. Other research teams have instead employed viral vector approaches for the long-term CNS-specific expression of BDNF or other neurotrophins. An adeno-associated viral (AAV) vector encoding Bdnf cDNA under the control of an astrocytic promoter, when injected into the striatum of R6/2 mice, was shown to delay the onset of HD-specific behavioral abnormalities. This effect may be related to increased neostriatal neurogenesis rather than the reduction of MSN death, as BDNF overexpression through AAV approaches has been shown to increase doublecortin immunoreactivity in R6/2 mice [215]. Other neurotrophins, when overexpressed, may support or potentiate the activity of BDNF. In the quinolinic acid rodent model of HD, rats that underwent the unilateral intrastriatal injection of a BDNF or Glial-derived neurotrophic factor (GDNF) AAV vector were protected against the loss of critical interneuron subpopulations affected early in this model [216]. It is important to note that, aside from translational issues concerning invasiveness and the fine control of gene expression in vivo, direct BDNF synthesis via viral vector approaches may also exert harmful neural effects in humans that are not observed in models with short lifespans. For example, it is unlikely that impairments of BDNF transport or metabolism will be directly rescued by BDNF administration alone. Increased and sustained activation of p75NTR in the context of reduced TrkB expression may instead exacerbate disease-specific outcomes in HD patients and disease models. For these reasons, these “direct” strategies have largely failed to gain practical success outside of preclinical models.
Engineered mesenchymal stem cell (MSC) transplant is under growing consideration as a therapeutic agent in HD for their neural repair properties, which are partially connected with their constitutive ability to release BDNF. At least three groups have reported that the intravenous treatment of MSC-like human dental pulp stem cells possesses the ability to alleviate key HD-like pathological features in the 3-nitropropionic (3-NP) rat HD model [217,218,219]. One of these groups illustrated increased BDNF immunoreactivity in the striatum of these 3-NP/MSC treated rats vs. controls [219]. Other groups described the mitigation of striatal atrophy and an extended lifespan in two additional HD models (YAC128 and R6/2) that were engrafted with intrastriatal human MSCs, although this study required the use of mice on immunocompromised backgrounds [220]. Nonetheless, Cellavita, a MSC-based therapy involving the regular systemic administration of human MSCs, is currently in a phase II clinical trial in Brazil for HD (NCT03252535). More sophisticated transplant and genetic expression strategies may prove to be more beneficial. The direct transplant of human neural stem cells into the striata of R6/2 and HdhQ140 mice resulted in improved motor and electrophysiological deficits in these models, thought to be facilitated by increased BDNF production [221]. One report, involving the engraftment of conditionally immortalized neural stem cell lines CTX0E03, documented increased counts of DARPP32- and BDNF-positive cells in the striatum of the quinolinic acid (QA) rat HD model. These cells appeared to have beneficial effects on inflammation markers on neighboring neurons, suspected to be related to BDNF [222]. Another report found that the intracerebral transplantation of BDNF-overexpressing human neural stem cells into QA-lesioned rats reduced the striatal expression of molecular markers of inflammation [223]. It is worth noting that the use of MSCs and other stem cells in transplantation-based approaches comes with its own sets of limitations. These include delivery concerns, tumorigenic potential, viability in vivo over extended timeframes, immunogenicity, and heterogeneity of these cells [224,225].
More indirect and less-invasive experimental manipulations that serve to broadly up-regulate BDNF may alleviate disease progression and offer immediate insight for the clinical management of HD in humans. Exercise enrichment, a modification which is well-connected to a beneficial CNS-wide BDNF response, was sufficient to rescue the loss of BDNF in the hippocampus and striatum of R6/1 mice [226]. Accordingly, the recent ENGAGE-HD (ISRCTN65378754) and PACE-HD (NCT03344601) clinical trials employed physical activity interventions in HD individuals but did not evaluate BDNF expression as primary or secondary measures of interest. Unfortunately, while non-control participants reported greater levels of physical activity and fitness, these studies reported no effect for activity on the UHDRS motor score. It is important to note that physical activity interventions may only be possible in pre-manifest and early manifest stages due to the profound loss of movement control in HD, thereby limiting their therapeutic potential. Environmental modification and enrichment (EME) also appear to bolster BDNF expression in preclinical models. Frequent introduction of novel objects into the home environments of R6/1 strongly upregulated cortical BDNF expression and delayed the onset of motor symptoms in this model [226]. Intriguingly, physical-activity and EME appeared to differentially modify BDNF localization in an exon-specific manner in R6/1 mice, pointing to discrete mechanisms that have a shared outcome of enhanced BDNF signaling [227]. Other EME interventions in HD individuals such as music, language, and multidisciplinary/hybrid tasks have suggested modest benefits for these forms of cognitive and social stimulation [228,229,230,231]. As above, results of human HD trials involving these and other environmental modifiers are mixed, and it is still unclear if the effects of these limited interventions are related to increased BDNF expression in the CNS. Ultimately, future management of HD will likely involve combinations of physical activity and other interventions alongside emerging pharmacotherapeutics.

5.2. BDNF-Targeting Therapeutics

The repurposing of emerging small molecule compounds, especially those aimed at alleviating impairment in key cellular pathways, may also indirectly up-regulate BDNF or its signaling and therefore offer similar promise to more direct pharmacotherapies in HD. For example, at least two groups have evaluated the metabotropic glutamate receptor 2/3 (mGluR2/3) agonist LY379268 in preclinical HD models for its enigmatic BDNF-stimulating property [232,233]. Some of these reports confirmed the up-regulation of Bdnf transcription [232]. Other groups attempted to employ small molecules to indirectly supplement a BDNF-TrkB signaling tone otherwise lost in disease progression. The A2AR agonist inosine mitigated 3-NP-induced motor abnormalities and body weight loss by increasing BDNF signaling through the TrkB-ERK-CREB pathway [234]. Another study showed that the activation of sphingosine-1-phosphate receptor 5 (S1PR5) by the agonist A-971432 delayed the onset of motor dysfunction in R6/2 mice by the activation of Trk-downstream AKT and ERK pathways [235]. Other small molecules, including those with action at TrkB [236,237], p75NTR [238], and REST [239] have also shown limited beneficial effects in HD models, likely due in part to the partial rescue of BDNF expression or TrkB activity through their pharmacological targets.
The use of the sigma-1 receptor (S1R) agonist pridopidine is under increasing consideration for the treatment of HD. Pridopidine was shown to promote a significant increase in BDNF axonal trafficking, TrkB signaling at the synapse, and enhanced release of glutamate, albeit within a reconstituted “brain-on-a-chip” model of HD [44]. A bulk RNA-seq approach in the YAC128 model also indicated the rescue of neurotrophin expression associated with pridopidine treatment [240]. It is still unknown by which mechanism(s) the activation of the sigma-1 receptor results in the restored trafficking of BDNF, but it is suggested that the reduction of ER stress and/or stimulation of proteolytic degradation of mHTT aggregates may have beneficial knock-on effects on axonal transport or expression in this model [241,242]. Nevertheless, the use of pridopidine in HD has advanced into HD-focused clinical trials. Of note, the PRIDE-HD phase II exploratory trial (NCT02006472) documented a delay in the reduction in total functional capacity in HD patients over one year of pridopidine treatment, but is not known if BDNF up-regulation was responsible for this outcome [243,244]. Pridopidine is currently in a multicenter phase III clinical trial for patients with early-stage HD; however, this trial will not evaluate BDNF expression nor non-neuropsychological outcomes as part of its aims (NCT04556656). Another indirect pharmacological strategy, inhibition of phosphodiesterase 4 (PDE4), was also proposed as a therapeutic approach for its ability to stimulate neural BDNF expression. To date, two phase I clinical trials have taken place that evaluated safety and tolerability of the PDE4 small molecule inhibitor GSK356278, with the eventual goal of use of this compound in HD (NCT01602900, NCT01573819). It is not known whether GSK356278 provoked an increase of CSF BDNF levels in these studies, and it has yet to advance to phase II trials for use in HD.
Still, other research terms have engineered novel peptidergic compounds that specifically modify BDNF expression. Two intriguing reports from Couly et al. indicated that a 23 amino acid peptide “P42” constructed from HTT, when delivered systemically over several weeks, was sufficient to rescue deficits in Schaffer collateral LTP and had modest effects in restoring motor function in the R6/2 model. These effects were suggested to be related to the up-regulation of BDNF and increasing activity at TrkB in P42-treated mice, although it is still unclear the set of cellular mechanisms by which this effect was produced [245]. Somewhat encouragingly, the positive effects of P42 are potentiated by habitat enrichment in this model, an environmental modification that is well-connected with a beneficial BDNF response [246]. Another group reported that the intranasal administration of a recombinant vasoactive intestinal peptide (VIP) upregulated striatal BDNF in the R6/1 model, likely through coupling with the cAMP–PKA–CREB pathway (Figure 2) [247]. In summary, advancements in neurotrophin drug discovery and delivery will unlock previously inaccessible therapeutic options for the treatment of HD and various neurodegenerative diseases.

6. Discussion

The dysregulation of BDNF is a well-documented and significant molecular alteration in HD and plays a critical role in its pathogenesis. Substantial evidence suggests that reduced BDNF trophic support and impaired TrkB signaling underlie the degeneration of striatal projection neurons, the most vulnerable cell type in HD. A deeper understanding of the mechanisms leading to deficits in BDNF-TrkB signaling is essential for developing effective neuroprotective treatments for HD.
Despite continued research into neuron-protective factors for use in neurodegenerative diseases, neurotrophin-based therapeutic approaches have yet to gain serious attention in clinical practice. The set of challenges restraining the identification and development of CNS pharmacotherapeutics is exemplified by BDNF. These practical hurdles include poor half-live in vivo, off-target effects, and poor blood–brain barrier (BBB) penetrability. Drugs that aim to target BDNF expression in vivo must also overcome these obstacles. Still, there currently appears to be more promise for neurotrophin-based therapies in HD than in recent history, owing to advances in the understanding of BBB physiology. Improvements in CNS drug delivery methods, including transcranial focused ultrasound-induced permeabilization [248,249], nasal atomization and insufflation [250,251], and novel drug carriers such as immunoliposomes [252] and nanoparticles (NCT03815916, NCT03020017), have all recently shown positive results in humans for CNS disorders or in animal studies. Continued advances in CNS drug delivery will therefore always be of critical importance in HD, the neurotrophin field, and neurodegenerative research more broadly.
Due to the complexities involved with neurotrophin drug delivery, and the monogenic nature of HD, many compounds investigated within the HD clinical trial pipeline to date did not target BDNF or TrkB directly, but instead acted on mHTT expression itself. It is suspected, but not yet established, that the downregulation of mHTT or reduction of mHTT aggregation through RNA interference may indirectly increase BDNF expression in neurons and thereby provide benefits to neuronal health [130]. The administration of antisense oligonucleotides (ASOs) against HTT or mHTT transcripts therefore remains a popular therapeutic approach. Recent and ongoing ASO clinical trials undertaken by Ionis Pharmaceuticals (NCT02519036, NCT03761849, NCT05686551) and Wave Life Sciences (NCT03225846, NCT03225833, NCT05032196) unfortunately have shown modest or no improvement in measures of disease progression. Similar discouraging results from repurposed oral HTT-splicing modulators, such as Branaplam, were recently reported (NCT05111249). It is unclear if CSF or serum BDNF were evaluated within these trials, making comparisons with preclinical research in this direction difficult. It is also important to note that some studies have suggested that the presence of mHTT inclusions is instead a neuronal resiliency mechanism which serves to impede other toxic gain-of-functions of mHTT [253]. HTT-lowering strategies may benefit from the more precise and mutant allele-specific inhibition of HTT targeting, rather than the broad silencing of wild-type and mutant HTT transcripts, given the growing repertoire of observed wild-type HTT neural functions related to BDNF. The continued improvement in allele-selective ASO design and delivery will undoubtedly have considerable implications for HD.
In the present review, we illustrated numerous ways in which BDNF expression, transport, and release are compromised in HD. However, there remains much to be uncovered in relation to the intersection of BDNF dysregulation and HD. For example, questions revolving around the ability of BDNF to modulate CNS inflammation associated with HD remain underinvestigated. The nature of the glial inflammatory response in the presence of reduced BDNF in HD may be an especially important research direction because some neurotrophins, including BDNF, appear to temper the glial inflammatory response to favor anti-inflammatory modes of activation in certain models [254,255,256]. The broad use of BDNF-directed therapies may also have undesired effects on other cell types and within different neuroanatomical substrates, which provides an incentive for the selective regional targeting of BDNF-based treatments. The development of therapeutics and future clinical management of HD will therefore benefit from considering the effect of BDNF loss on diverse neural and non-neural cellular functions. Ultimately, despite stubborn challenges to their use in the CNS, BDNF-directed pharmacotherapies may represent a desperately needed disease-modifying treatment for HD.

7. Conclusions

HD remains an incredibly challenging neurodegenerative disease to address because of the paucity of available molecular targets and lack of suitable therapeutic intervention strategies. Although a few FDA-approved pharmacotherapies exist to manage the classic choreic symptoms coupled with HD, none are able to slow or prevent HD pathogenesis. Therefore, candidate approaches involving broadly protective factors may prove to be crucial disease-modifying treatments for HD individuals. BDNF is a neurotrophin that is critically intertwined with an array of vital neuronal signaling pathways through TrkB. BDNF-TrkB signaling is strongly supportive of neuronal resilience, longevity, and normal function. The loss of BDNF in HD therefore provides both an explanation for the array of molecular, functional, and cognitive outcomes that manifest in HD as well as a potential pathway for treatment. Here, we extensively discussed multiple major HD mechanisms that have a shared outcome of neural BDNF reduction. We provided a special focus on the epigenetic and transcriptional dysregulation of BDNF in HD, a critical characteristic of its early molecular pathogenesis. Finally, we offered a critical analysis of emerging therapeutic approaches involving this neurotrophin within HD. Despite the promise of neurotrophin-based therapeutic approaches, there unfortunately remain major practical obstacles for their future use. These include delivery to deep neuroanatomical structures, poor pharmacokinetic profile, and uncertain efficacy in humans. Emerging viral-based strategies, Trk agonists, and brain-penetrant small molecules or peptides may soon overcome these challenges. Ultimately, we propose that neurotrophin therapies will likely form part of a multi-pronged series of interventions for HD involving lifestyle modification, symptomatic treatment, and small molecule-based approaches to modify disease progression.

8. Search Methods and Strategy

Sources included in this literature review were populated from three major literature databases: Pubmed, Web of Science (Thompson Routers), and Scopus. These databases were supplemented from multiple gray literature sources, including Google Scholar, clinicaltrials.gov, EU Clinical Trials Register, and the bioRxiv preprint server. Our search strategy involved Boolean strings constructed from MeSH terms (e.g., “Brain-Derived Neurotrophic Factor” [Mesh] AND “Huntingtin Protein” [Mesh] AND “Axonal Transport” [Mesh]). These search strings were complemented by plain text searching within the MeSH string results, the examination of cross-referenced sources, and personal knowledge of relevant articles. Each of the authors independently executed these literature searches and prioritized articles that were published in the previous 5 years. There were no a priori inclusion or exclusion criteria.

Author Contributions

A.S., N.B.A. and H.Y. contributed to the conceptualization, writing, review, and editing of this review article. N.B.A. constructed the figures. All authors have read and agreed to the published version of the manuscript.

Funding

The writing of this article was supported through NIH R01NS111014 (H.Y.).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors jointly declare no conflict of interest.

References

  1. Fisher, E.R.; Hayden, M.R. Multisource ascertainment of Huntington disease in Canada: Prevalence and population at risk. Mov. Disord. 2014, 29, 105–114. [Google Scholar] [CrossRef]
  2. George Huntington, M.D. On chorea. J. Neuropsychiatry Clin. Neurosci. 2003, 15, 109–112. [Google Scholar] [CrossRef]
  3. The Huntington’s Disease Collaborative Research Group. A Novel Gene Containing a Trinucleotide Repeat That Is Expanded and Unstable on Huntingtons-Disease Chromosomes. Cell 1993, 72, 971–983. [Google Scholar] [CrossRef]
  4. Wright, G.E.B.; Black, H.F.; Collins, J.A.; Gall-Duncan, T.; Caron, N.S.; Pearson, C.E.; Hayden, M.R. Interrupting sequence variants and age of onset in Huntington’s disease: Clinical implications and emerging therapies. Lancet Neurol. 2020, 19, 930–939. [Google Scholar] [CrossRef]
  5. McDonnell, E.I.; Wang, Y.; Goldman, J.; Marder, K. Age of Onset of Huntington’s Disease in Carriers of Reduced Penetrance Alleles. Mov. Disord. 2021, 36, 2958–2961. [Google Scholar] [CrossRef]
  6. Wright, G.E.B.; Collins, J.A.; Kay, C.; McDonald, C.; Dolzhenko, E.; Xia, Q.; Becanovic, K.; Drogemoller, B.I.; Semaka, A.; Nguyen, C.M.; et al. Length of Uninterrupted CAG, Independent of Polyglutamine Size, Results in Increased Somatic Instability, Hastening Onset of Huntington Disease. Am. J. Hum. Genet. 2019, 104, 1116–1126. [Google Scholar] [CrossRef]
  7. Tabrizi, S.J.; Flower, M.D.; Ross, C.A.; Wild, E.J. Huntington disease: New insights into molecular pathogenesis and therapeutic opportunities. Nat. Rev. Neurol. 2020, 16, 529–546. [Google Scholar] [CrossRef]
  8. Genetic Modifiers of Huntington’s Disease Consortium. CAG Repeat Not Polyglutamine Length Determines Timing of Huntington’s Disease Onset. Cell 2019, 178, 887–900. [Google Scholar] [CrossRef]
  9. Khristich, A.N.; Mirkin, S.M. On the wrong DNA track: Molecular mechanisms of repeat-mediated genome instability. J. Biol. Chem. 2020, 295, 4134–4170. [Google Scholar] [CrossRef]
  10. Swami, M.; Hendricks, A.E.; Gillis, T.; Massood, T.; Mysore, J.; Myers, R.H.; Wheeler, V.C. Somatic expansion of the Huntington’s disease CAG repeat in the brain is associated with an earlier age of disease onset. Hum. Mol. Genet. 2009, 18, 3039–3047. [Google Scholar] [CrossRef]
  11. Loupe, J.M.; Pinto, R.M.; Kim, K.H.; Gillis, T.; Mysore, J.S.; Andrew, M.A.; Kovalenko, M.; Murtha, R.; Seong, I.; Gusella, J.F.; et al. Promotion of somatic CAG repeat expansion by Fan1 knock-out in Huntington’s disease knock-in mice is blocked by Mlh1 knock-out. Hum. Mol. Genet. 2020, 29, 3044–3053. [Google Scholar] [CrossRef]
  12. Monckton, D.G. The Contribution of Somatic Expansion of the CAG Repeat to Symptomatic Development in Huntington’s Disease: A Historical Perspective. J. Huntington’s Dis. 2021, 10, 7–33. [Google Scholar] [CrossRef]
  13. Kacher, R.; Lejeune, F.X.; Noel, S.; Cazeneuve, C.; Brice, A.; Humbert, S.; Durr, A. Propensity for somatic expansion increases over the course of life in Huntington disease. eLife 2021, 10, e64674. [Google Scholar] [CrossRef]
  14. Reiner, A.; Albin, R.L.; Anderson, K.D.; D’Amato, C.J.; Penney, J.B.; Young, A.B. Differential loss of striatal projection neurons in Huntington disease. Proc. Natl. Acad. Sci. USA 1988, 85, 5733–5737. [Google Scholar] [CrossRef]
  15. Deng, Y.P.; Albin, R.L.; Penney, J.B.; Young, A.B.; Anderson, K.D.; Reiner, A. Differential loss of striatal projection systems in Huntington’s disease: A quantitative immunohistochemical study. J. Chem. Neuroanat. 2004, 27, 143–164. [Google Scholar] [CrossRef]
  16. Matsushima, A.; Pineda, S.S.; Crittenden, J.R.; Lee, H.; Galani, K.; Mantero, J.; Tombaugh, G.; Kellis, M.; Heiman, M.; Graybiel, A.M. Transcriptional vulnerabilities of striatal neurons in human and rodent models of Huntington’s disease. Nat. Commun. 2023, 14, 282. [Google Scholar] [CrossRef]
  17. Lee, H.; Fenster, R.J.; Pineda, S.S.; Gibbs, W.S.; Mohammadi, S.; Davila-Velderrain, J.; Garcia, F.J.; Therrien, M.; Novis, H.S.; Gao, F.; et al. Cell Type-Specific Transcriptomics Reveals that Mutant Huntingtin Leads to Mitochondrial RNA Release and Neuronal Innate Immune Activation. Neuron 2020, 107, 891–908. [Google Scholar] [CrossRef]
  18. Mätlik, K.; Baffuto, M.; Kus, L.; Davis, D.A.; Paul, M.R.; Carroll, T.S.; Heintz, N. Cell Type Specific CAG Repeat Expansion and Toxicity of Mutant Huntingtin in Human Striatal and Cerebellar Cell Types. bioRxiv 2023. [Google Scholar] [CrossRef]
  19. Vonsattel, J.P.; Myers, R.H.; Stevens, T.J.; Ferrante, R.J.; Bird, E.D.; Richardson, E.P., Jr. Neuropathological classification of Huntington’s disease. J. Neuropathol. Exp. Neurol. 1985, 44, 559–577. [Google Scholar] [CrossRef]
  20. Vonsattel, J.P.; Keller, C.; Del Pilar Amaya, M. Neuropathology of Huntington’s disease. Handb. Clin. Neurol. 2008, 89, 599–618. [Google Scholar] [CrossRef]
  21. Singh-Bains, M.K.; Mehrabi, N.F.; Sehji, T.; Austria, M.D.R.; Tan, A.Y.S.; Tippett, L.J.; Dragunow, M.; Waldvogel, H.J.; Faull, R.L.M. Cerebellar degeneration correlates with motor symptoms in Huntington disease. Ann. Neurol. 2019, 85, 396–405. [Google Scholar] [CrossRef]
  22. Weiss, A.; Liguore, W.; Brandon, K.; Wang, X.; Liu, Z.; Kroenke, C.; McBride, J. Alterations of fractional anisotropy throughout cortico-basal ganglia gray matter in a macaque model of Huntington’s Disease. Curr. Res. Neurobiol. 2023, 4, 100090. [Google Scholar] [CrossRef]
  23. Espinoza, F.A.; Turner, J.A.; Vergara, V.M.; Miller, R.L.; Mennigen, E.; Liu, J.; Misiura, M.B.; Ciarochi, J.; Johnson, H.J.; Long, J.D.; et al. Whole-Brain Connectivity in a Large Study of Huntington’s Disease Gene Mutation Carriers and Healthy Controls. Brain Connect. 2018, 8, 166–178. [Google Scholar] [CrossRef]
  24. Gregory, S.; Johnson, E.; Byrne, L.M.; Rodrigues, F.B.; Henderson, A.; Moss, J.; Thomas, D.; Zhang, H.; De Vita, E.; Tabrizi, S.J.; et al. Characterizing White Matter in Huntington’s Disease. Mov. Disord. Clin. Pract. 2020, 7, 52–60. [Google Scholar] [CrossRef]
  25. Zhang, J.; Gregory, S.; Scahill, R.I.; Durr, A.; Thomas, D.L.; Lehericy, S.; Rees, G.; Tabrizi, S.J.; Zhang, H.; TrackOn-HD Investigators. In vivo characterization of white matter pathology in premanifest huntington’s disease. Ann. Neurol. 2018, 84, 497–504. [Google Scholar] [CrossRef]
  26. van Duijn, E.; Kingma, E.M.; Timman, R.; Zitman, F.G.; Tibben, A.; Roos, R.A.; van der Mast, R.C. Cross-sectional study on prevalences of psychiatric disorders in mutation carriers of Huntington’s disease compared with mutation-negative first-degree relatives. J. Clin. Psychiatry 2008, 69, 1804–1810. [Google Scholar] [CrossRef]
  27. Pierce, J.E.; Peron, J. The basal ganglia and the cerebellum in human emotion. Soc. Cogn. Affect. Neurosci. 2020, 15, 599–613. [Google Scholar] [CrossRef]
  28. Guo, Y.; Schmitz, T.W.; Mur, M.; Ferreira, C.S.; Anderson, M.C. A supramodal role of the basal ganglia in memory and motor inhibition: Meta-analytic evidence. Neuropsychologia 2018, 108, 117–134. [Google Scholar] [CrossRef]
  29. Ahveninen, L.M.; Stout, J.C.; Georgiou-Karistianis, N.; Lorenzetti, V.; Glikmann-Johnston, Y. Reduced amygdala volumes are related to motor and cognitive signs in Huntington’s disease: The IMAGE-HD study. NeuroImage Clin. 2018, 18, 881–887. [Google Scholar] [CrossRef]
  30. Petrella, L.I.; Castelhano, J.M.; Ribeiro, M.; Sereno, J.V.; Goncalves, S.I.; Laco, M.N.; Hayden, M.R.; Rego, A.C.; Castelo-Branco, M. A whole brain longitudinal study in the YAC128 mouse model of Huntington’s disease shows distinct trajectories of neurochemical, structural connectivity and volumetric changes. Hum. Mol. Genet. 2018, 27, 2125–2137. [Google Scholar] [CrossRef]
  31. Coppen, E.M.; Jacobs, M.; van den Berg-Huysmans, A.A.; van der Grond, J.; Roos, R.A.C. Grey matter volume loss is associated with specific clinical motor signs in Huntington’s disease. Park. Relat. Disord. 2018, 46, 56–61. [Google Scholar] [CrossRef]
  32. Zhang, H.; Zhang, C.; Vincent, J.; Zala, D.; Benstaali, C.; Sainlos, M.; Grillo-Bosch, D.; Daburon, S.; Coussen, F.; Cho, Y.; et al. Modulation of AMPA receptor surface diffusion restores hippocampal plasticity and memory in Huntington’s disease models. Nat. Commun. 2018, 9, 4272. [Google Scholar] [CrossRef]
  33. Glikmann-Johnston, Y.; Carmichael, A.M.; Mercieca, E.C.; Stout, J.C. ‘Real-life’ hippocampal-dependent spatial memory impairments in Huntington’s disease. Cortex 2019, 119, 46–60. [Google Scholar] [CrossRef]
  34. Glikmann-Johnston, Y.; Mercieca, E.C.; Carmichael, A.M.; Alexander, B.; Harding, I.H.; Stout, J.C. Hippocampal and striatal volumes correlate with spatial memory impairment in Huntington’s disease. J. Neurosci. Res. 2021, 99, 2948–2963. [Google Scholar] [CrossRef]
  35. Ellis, N.; Tee, A.; McAllister, B.; Massey, T.; McLauchlan, D.; Stone, T.; Correia, K.; Loupe, J.; Kim, K.H.; Barker, D.; et al. Genetic Risk Underlying Psychiatric and Cognitive Symptoms in Huntington’s Disease. Biol. Psychiatry 2020, 87, 857–865. [Google Scholar] [CrossRef]
  36. McAllister, B.; Gusella, J.F.; Landwehrmeyer, G.B.; Lee, J.M.; MacDonald, M.E.; Orth, M.; Rosser, A.E.; Williams, N.M.; Holmans, P.; Jones, L.; et al. Timing and Impact of Psychiatric, Cognitive, and Motor Abnormalities in Huntington Disease. Neurology 2021, 96, e2395–e2406. [Google Scholar] [CrossRef]
  37. Kachian, Z.R.; Cohen-Zimerman, S.; Bega, D.; Gordon, B.; Grafman, J. Suicidal ideation and behavior in Huntington’s disease: Systematic review and recommendations. J. Affect. Disord. 2019, 250, 319–329. [Google Scholar] [CrossRef]
  38. Achenbach, J.; Saft, C. Another Perspective on Huntington’s Disease: Disease Burden in Family Members and Pre-Manifest HD when Compared to Genotype-Negative Participants from ENROLL-HD. Brain Sci. 2021, 11, 1621. [Google Scholar] [CrossRef]
  39. Seefelder, M.; Kochanek, S. A meta-analysis of transcriptomic profiles of Huntington’s disease patients. PLoS ONE 2021, 16, e0253037. [Google Scholar] [CrossRef]
  40. Creus-Muncunill, J.; Badillos-Rodriguez, R.; Garcia-Forn, M.; Masana, M.; Garcia-Diaz Barriga, G.; Guisado-Corcoll, A.; Alberch, J.; Malagelada, C.; Delgado-Garcia, J.M.; Gruart, A.; et al. Increased translation as a novel pathogenic mechanism in Huntington’s disease. Brain 2019, 142, 3158–3175. [Google Scholar] [CrossRef]
  41. Pircs, K.; Drouin-Ouellet, J.; Horvath, V.; Gil, J.; Rezeli, M.; Garza, R.; Grassi, D.A.; Sharma, Y.; St-Amour, I.; Harris, K.; et al. Distinct subcellular autophagy impairments in induced neurons from patients with Huntington’s disease. Brain 2022, 145, 3035–3057. [Google Scholar] [CrossRef]
  42. Oh, Y.M.; Lee, S.W.; Kim, W.K.; Chen, S.; Church, V.A.; Cates, K.; Li, T.; Zhang, B.; Dolle, R.E.; Dahiya, S.; et al. Age-related Huntington’s disease progression modeled in directly reprogrammed patient-derived striatal neurons highlights impaired autophagy. Nat. Neurosci. 2022, 25, 1420–1433. [Google Scholar] [CrossRef]
  43. Agrawal, S.; Fox, J.H. Novel proteomic changes in brain mitochondria provide insights into mitochondrial dysfunction in mouse models of Huntington’s disease. Mitochondrion 2019, 47, 318–329. [Google Scholar] [CrossRef]
  44. Lenoir, S.; Lahaye, R.A.; Vitet, H.; Scaramuzzino, C.; Virlogeux, A.; Capellano, L.; Genoux, A.; Gershoni-Emek, N.; Geva, M.; Hayden, M.R.; et al. Pridopidine rescues BDNF/TrkB trafficking dynamics and synapse homeostasis in a Huntington disease brain-on-a-chip model. Neurobiol. Dis. 2022, 173, 105857. [Google Scholar] [CrossRef]
  45. Chen, S.; Yu, C.; Rong, L.; Li, C.H.; Qin, X.; Ryu, H.; Park, H. Altered Synaptic Vesicle Release and Ca2+ Influx at Single Presynaptic Terminals of Cortical Neurons in a Knock-in Mouse Model of Huntington’s Disease. Front. Mol. Neurosci. 2018, 11, 478. [Google Scholar] [CrossRef]
  46. Colucci-D’Amato, L.; Speranza, L.; Volpicelli, F. Neurotrophic Factor BDNF, Physiological Functions and Therapeutic Potential in Depression, Neurodegeneration and Brain Cancer. Int. J. Mol. Sci. 2020, 21, 7777. [Google Scholar] [CrossRef]
  47. Ibrahim, A.M.; Chauhan, L.; Bhardwaj, A.; Sharma, A.; Fayaz, F.; Kumar, B.; Alhashmi, M.; AlHajri, N.; Alam, M.S.; Pottoo, F.H. Brain-Derived Neurotropic Factor in Neurodegenerative Disorders. Biomedicines 2022, 10, 1143. [Google Scholar] [CrossRef]
  48. Jeanneteau, F.; Arango-Lievano, M.; Chao, M. Neurotrophin and Synaptogenesis. In Synapse Development and Maturation; Academic Press: Cambridge, MA, USA, 2020; pp. 167–192. [Google Scholar]
  49. Pugazhenthi, S.; Wang, M.; Pham, S.; Sze, C.I.; Eckman, C.B. Downregulation of CREB expression in Alzheimer’s brain and in Abeta-treated rat hippocampal neurons. Mol. Neurodegener. 2011, 6, 60. [Google Scholar] [CrossRef]
  50. Nguyen, K.Q.; Rymar, V.V.; Sadikot, A.F. Impaired TrkB Signaling Underlies Reduced BDNF-Mediated Trophic Support of Striatal Neurons in the R6/2 Mouse Model of Huntington’s Disease. Front. Cell Neurosci. 2016, 10, 37. [Google Scholar] [CrossRef]
  51. Forlenza, O.V.; Diniz, B.S.; Teixeira, A.L.; Radanovic, M.; Talib, L.L.; Rocha, N.P.; Gattaz, W.F. Lower Cerebrospinal Fluid Concentration of Brain-Derived Neurotrophic Factor Predicts Progression from Mild Cognitive Impairment to Alzheimer’s Disease. Neuromol. Med. 2015, 17, 326–332. [Google Scholar] [CrossRef]
  52. Caron, N.S.; Haqqani, A.S.; Sandhu, A.; Aly, A.E.; Findlay Black, H.; Bone, J.N.; McBride, J.L.; Abulrob, A.; Stanimirovic, D.; Leavitt, B.R.; et al. Cerebrospinal fluid biomarkers for assessing Huntington disease onset and severity. Brain Commun. 2022, 4, fcac309. [Google Scholar] [CrossRef] [PubMed]
  53. Abassi, M.; Morawski, B.M.; Nakigozi, G.; Nakasujja, N.; Kong, X.; Meya, D.B.; Robertson, K.; Gray, R.; Wawer, M.J.; Sacktor, N.; et al. Cerebrospinal fluid biomarkers and HIV-associated neurocognitive disorders in HIV-infected individuals in Rakai, Uganda. J. Neurovirol. 2017, 23, 369–375. [Google Scholar] [CrossRef] [PubMed]
  54. Jiao, S.S.; Shen, L.L.; Zhu, C.; Bu, X.L.; Liu, Y.H.; Liu, C.H.; Yao, X.Q.; Zhang, L.L.; Zhou, H.D.; Walker, D.G.; et al. Brain-derived neurotrophic factor protects against tau-related neurodegeneration of Alzheimer’s disease. Transl. Psychiatry 2016, 6, e907. [Google Scholar] [CrossRef]
  55. Bachis, A.; Avdoshina, V.; Zecca, L.; Parsadanian, M.; Mocchetti, I. Human immunodeficiency virus type 1 alters brain-derived neurotrophic factor processing in neurons. J. Neurosci. 2012, 32, 9477–9484. [Google Scholar] [CrossRef] [PubMed]
  56. Almeida, M.F.; Chaves, R.S.; Silva, C.M.; Chaves, J.C.S.; Melo, K.P.; Ferrari, M.F.R. BDNF trafficking and signaling impairment during early neurodegeneration is prevented by moderate physical activity. IBRO Rep. 2016, 1, 19–31. [Google Scholar] [CrossRef]
  57. Hampel, H.; Mesulam, M.M.; Cuello, A.C.; Khachaturian, A.S.; Vergallo, A.; Farlow, M.R.; Snyder, P.J.; Giacobini, E.; Khachaturian, Z.S. Revisiting the Cholinergic Hypothesis in Alzheimer’s Disease: Emerging Evidence from Translational and Clinical Research. J. Prev. Alzheimers Dis. 2019, 6, 2–15. [Google Scholar] [CrossRef]
  58. Moya-Alvarado, G.; Tiburcio-Felix, R.; Ibanez, M.R.; Aguirre-Soto, A.A.; Guerra, M.V.; Wu, C.; Mobley, W.C.; Perlson, E.; Bronfman, F.C. BDNF/TrkB signaling endosomes in axons coordinate CREB/mTOR activation and protein synthesis in the cell body to induce dendritic growth in cortical neurons. eLife 2023, 12, e77455. [Google Scholar] [CrossRef]
  59. Guo, W.; Nagappan, G.; Lu, B. Differential effects of transient and sustained activation of BDNF-TrkB signaling. Dev. Neurobiol. 2018, 78, 647–659. [Google Scholar] [CrossRef]
  60. Liu, Y.; Ma, C.; Li, H.; Shen, H.; Li, X.; Fu, X.; Wu, J.; Chen, G. Nogo-A/Pir-B/TrkB Signaling Pathway Activation Inhibits Neuronal Survival and Axonal Regeneration after Experimental Intracerebral Hemorrhage in Rats. J. Mol. Neurosci. 2019, 69, 360–370. [Google Scholar] [CrossRef]
  61. Arenas, Y.M.; Martinez-Garcia, M.; Llansola, M.; Felipo, V. Enhanced BDNF and TrkB Activation Enhance GABA Neurotransmission in Cerebellum in Hyperammonemia. Int. J. Mol. Sci. 2022, 23, 11770. [Google Scholar] [CrossRef]
  62. Hempstead, B.L.; Martin-Zanca, D.; Kaplan, D.R.; Parada, L.F.; Chao, M.V. High-affinity NGF binding requires coexpression of the trk proto-oncogene and the low-affinity NGF receptor. Nature 1991, 350, 678–683. [Google Scholar] [CrossRef]
  63. Greenwood, S.G.; Montroull, L.; Volosin, M.; Scharfman, H.E.; Teng, K.K.; Light, M.; Torkin, R.; Maxfield, F.; Hempstead, B.L.; Friedman, W.J. A Novel Neuroprotective Mechanism for Lithium That Prevents Association of the p75(NTR)-Sortilin Receptor Complex and Attenuates proNGF-Induced Neuronal Death In Vitro and In Vivo. eNeuro 2018, 5, 1–14. [Google Scholar] [CrossRef] [PubMed]
  64. Speidell, A.; Asuni, G.P.; Wakulski, R.; Mocchetti, I. Up-regulation of the p75 neurotrophin receptor is an essential mechanism for HIV-gp120 mediated synaptic loss in the striatum. Brain Behav. Immun. 2020, 89, 371–379. [Google Scholar] [CrossRef] [PubMed]
  65. Ginsberg, S.D.; Che, S.; Wuu, J.; Counts, S.E.; Mufson, E.J. Down regulation of trk but not p75NTR gene expression in single cholinergic basal forebrain neurons mark the progression of Alzheimer’s disease. J. Neurochem. 2006, 97, 475–487. [Google Scholar] [CrossRef] [PubMed]
  66. Ferrer, I.; Marin, C.; Rey, M.J.; Ribalta, T.; Goutan, E.; Blanco, R.; Tolosa, E.; Marti, E. BDNF and full-length and truncated TrkB expression in Alzheimer disease. Implications in therapeutic strategies. J. Neuropathol. Exp. Neurol. 1999, 58, 729–739. [Google Scholar] [CrossRef]
  67. Parain, K.; Murer, M.G.; Yan, Q.; Faucheux, B.; Agid, Y.; Hirsch, E.; Raisman-Vozari, R. Reduced expression of brain-derived neurotrophic factor protein in Parkinson’s disease substantia nigra. Neuroreport 1999, 10, 557–561. [Google Scholar] [CrossRef]
  68. Szego, E.M.; Gerhardt, E.; Outeiro, T.F.; Kermer, P. Dopamine-depletion and increased alpha-synuclein load induce degeneration of cortical cholinergic fibers in mice. J. Neurol. Sci. 2011, 310, 90–95. [Google Scholar] [CrossRef]
  69. Kang, S.S.; Zhang, Z.; Liu, X.; Manfredsson, F.P.; Benskey, M.J.; Cao, X.; Xu, J.; Sun, Y.E.; Ye, K. TrkB neurotrophic activities are blocked by alpha-synuclein, triggering dopaminergic cell death in Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2017, 114, 10773–10778. [Google Scholar] [CrossRef]
  70. Danelon, V.; Montroull, L.E.; Unsain, N.; Barker, P.A.; Masco, D.H. Calpain-dependent truncated form of TrkB-FL increases in neurodegenerative processes. Mol. Cell. Neurosci. 2016, 75, 81–92. [Google Scholar] [CrossRef]
  71. Vidaurre, O.G.; Gascon, S.; Deogracias, R.; Sobrado, M.; Cuadrado, E.; Montaner, J.; Rodriguez-Pena, A.; Diaz-Guerra, M. Imbalance of neurotrophin receptor isoforms TrkB-FL/TrkB-T1 induces neuronal death in excitotoxicity. Cell Death Dis. 2012, 3, e256. [Google Scholar] [CrossRef]
  72. Yanpallewar, S.U.; Barrick, C.A.; Buckley, H.; Becker, J.; Tessarollo, L. Deletion of the BDNF truncated receptor TrkB.T1 delays disease onset in a mouse model of amyotrophic lateral sclerosis. PLoS ONE 2012, 7, e39946. [Google Scholar] [CrossRef]
  73. Meeker, R.; Williams, K. Dynamic nature of the p75 neurotrophin receptor in response to injury and disease. J. Neuroimmune Pharmacol. 2014, 9, 615–628. [Google Scholar] [CrossRef]
  74. Wang, M.; Xie, Y.; Qin, D. Proteolytic cleavage of proBDNF to mBDNF in neuropsychiatric and neurodegenerative diseases. Brain Res. Bull. 2021, 166, 172–184. [Google Scholar] [CrossRef]
  75. Gerenu, G.; Martisova, E.; Ferrero, H.; Carracedo, M.; Rantamaki, T.; Ramirez, M.J.; Gil-Bea, F.J. Modulation of BDNF cleavage by plasminogen-activator inhibitor-1 contributes to Alzheimer’s neuropathology and cognitive deficits. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 991–1001. [Google Scholar] [CrossRef]
  76. Brito, V.; Puigdellivol, M.; Giralt, A.; del Toro, D.; Alberch, J.; Gines, S. Imbalance of p75(NTR)/TrkB protein expression in Huntington’s disease: Implication for neuroprotective therapies. Cell Death Dis. 2013, 4, e595. [Google Scholar] [CrossRef]
  77. Ferrer, I.; Goutan, E.; Marin, C.; Rey, M.J.; Ribalta, T. Brain-derived neurotrophic factor in Huntington disease. Brain Res. 2000, 866, 257–261. [Google Scholar] [CrossRef] [PubMed]
  78. Zuccato, C.; Ciammola, A.; Rigamonti, D.; Leavitt, B.R.; Goffredo, D.; Conti, L.; MacDonald, M.E.; Friedlander, R.M.; Silani, V.; Hayden, M.R.; et al. Loss of huntingtin-mediated BDNF gene transcription in Huntington’s disease. Science 2001, 293, 493–498. [Google Scholar] [CrossRef] [PubMed]
  79. Gauthier, L.R.; Charrin, B.C.; Borrell-Pages, M.; Dompierre, J.P.; Rangone, H.; Cordelieres, F.P.; De Mey, J.; MacDonald, M.E.; Lessmann, V.; Humbert, S.; et al. Huntingtin controls neurotrophic support and survival of neurons by enhancing BDNF vesicular transport along microtubules. Cell 2004, 118, 127–138. [Google Scholar] [CrossRef]
  80. Altar, C.A.; Cai, N.; Bliven, T.; Juhasz, M.; Conner, J.M.; Acheson, A.L.; Lindsay, R.M.; Wiegand, S.J. Anterograde transport of brain-derived neurotrophic factor and its role in the brain. Nature 1997, 389, 856–860. [Google Scholar] [CrossRef] [PubMed]
  81. Baquet, Z.C.; Gorski, J.A.; Jones, K.R. Early striatal dendrite deficits followed by neuron loss with advanced age in the absence of anterograde cortical brain-derived neurotrophic factor. J. Neurosci. 2004, 24, 4250–4258. [Google Scholar] [CrossRef] [PubMed]
  82. Gines, S.; Bosch, M.; Marco, S.; Gavalda, N.; Diaz-Hernandez, M.; Lucas, J.J.; Canals, J.M.; Alberch, J. Reduced expression of the TrkB receptor in Huntington’s disease mouse models and in human brain. Eur. J. Neurosci. 2006, 23, 649–658. [Google Scholar] [CrossRef] [PubMed]
  83. Mai, M.; Akkad, A.D.; Wieczorek, S.; Saft, C.; Andrich, J.; Kraus, P.H.; Epplen, J.T.; Arning, L. No association between polymorphisms in the BDNF gene and age at onset in Huntington disease. BMC Med. Genet. 2006, 7, 79. [Google Scholar] [CrossRef]
  84. Alberch, J.; Lopez, M.; Badenas, C.; Carrasco, J.L.; Mila, M.; Munoz, E.; Canals, J.M. Association between BDNF Val66Met polymorphism and age at onset in Huntington disease. Neurology 2005, 65, 964–965. [Google Scholar] [CrossRef]
  85. Ciammola, A.; Sassone, J.; Cannella, M.; Calza, S.; Poletti, B.; Frati, L.; Squitieri, F.; Silani, V. Low brain-derived neurotrophic factor (BDNF) levels in serum of Huntington’s disease patients. Am. J. Med. Genet. Part B Neuropsychiatr. Genet. 2007, 144, 574–577. [Google Scholar] [CrossRef] [PubMed]
  86. Squitieri, F.; Orobello, S.; Cannella, M.; Martino, T.; Romanelli, P.; Giovacchini, G.; Frati, L.; Mansi, L.; Ciarmiello, A. Riluzole protects Huntington disease patients from brain glucose hypometabolism and grey matter volume loss and increases production of neurotrophins. Eur. J. Nucl. Med. Mol. Imaging 2009, 36, 1113–1120. [Google Scholar] [CrossRef] [PubMed]
  87. Zuccato, C.; Marullo, M.; Vitali, B.; Tarditi, A.; Mariotti, C.; Valenza, M.; Lahiri, N.; Wild, E.J.; Sassone, J.; Ciammola, A.; et al. Brain-derived neurotrophic factor in patients with Huntington’s disease. PLoS ONE 2011, 6, e22966. [Google Scholar] [CrossRef] [PubMed]
  88. Ou, Z.A.; Byrne, L.M.; Rodrigues, F.B.; Tortelli, R.; Johnson, E.B.; Foiani, M.S.; Arridge, M.; De Vita, E.; Scahill, R.I.; Heslegrave, A.; et al. Brain-derived neurotrophic factor in cerebrospinal fluid and plasma is not a biomarker for Huntington’s disease. Sci. Rep. 2021, 11, 3481. [Google Scholar] [CrossRef]
  89. Rodrigues, F.B.; Byrne, L.; McColgan, P.; Robertson, N.; Tabrizi, S.J.; Leavitt, B.R.; Zetterberg, H.; Wild, E.J. Cerebrospinal fluid total tau concentration predicts clinical phenotype in Huntington’s disease. J. Neurochem. 2016, 139, 22–25. [Google Scholar] [CrossRef]
  90. Wild, E.J.; Boggio, R.; Langbehn, D.; Robertson, N.; Haider, S.; Miller, J.R.; Zetterberg, H.; Leavitt, B.R.; Kuhn, R.; Tabrizi, S.J.; et al. Quantification of mutant huntingtin protein in cerebrospinal fluid from Huntington’s disease patients. J. Clin. Investig. 2015, 125, 1979–1986. [Google Scholar] [CrossRef]
  91. Byrne, L.M.; Rodrigues, F.B.; Blennow, K.; Durr, A.; Leavitt, B.R.; Roos, R.A.C.; Scahill, R.I.; Tabrizi, S.J.; Zetterberg, H.; Langbehn, D.; et al. Neurofilament light protein in blood as a potential biomarker of neurodegeneration in Huntington’s disease: A retrospective cohort analysis. Lancet Neurol. 2017, 16, 601–609. [Google Scholar] [CrossRef]
  92. Rodrigues, F.B.; Byrne, L.M.; Tortelli, R.; Johnson, E.B.; Wijeratne, P.A.; Arridge, M.; De Vita, E.; Ghazaleh, N.; Houghton, R.; Furby, H.; et al. Mutant huntingtin and neurofilament light have distinct longitudinal dynamics in Huntington’s disease. Sci. Transl. Med. 2020, 12, eabc2888. [Google Scholar] [CrossRef] [PubMed]
  93. Betti, L.; Palego, L.; Unti, E.; Mazzucchi, S.; Kiferle, L.; Palermo, G.; Bonuccelli, U.; Giannaccini, G.; Ceravolo, R. Brain-Derived Neurotrophic Factor (BDNF) and Serotonin Transporter (SERT) in Platelets of Patients with Mild Huntington’s Disease: Relationships with Social Cognition Symptoms. Arch. Ital. Biol. 2018, 156, 27–39. [Google Scholar] [CrossRef]
  94. Gutierrez, A.; Corey-Bloom, J.; Thomas, E.A.; Desplats, P. Evaluation of Biochemical and Epigenetic Measures of Peripheral Brain-Derived Neurotrophic Factor (BDNF) as a Biomarker in Huntington’s Disease Patients. Front. Mol. Neurosci. 2019, 12, 335. [Google Scholar] [CrossRef] [PubMed]
  95. Strand, A.D.; Baquet, Z.C.; Aragaki, A.K.; Holmans, P.; Yang, L.; Cleren, C.; Beal, M.F.; Jones, L.; Kooperberg, C.; Olson, J.M.; et al. Expression profiling of Huntington’s disease models suggests that brain-derived neurotrophic factor depletion plays a major role in striatal degeneration. J. Neurosci. 2007, 27, 11758–11768. [Google Scholar] [CrossRef]
  96. Paldino, E.; Giampa, C.; Montagna, E.; Angeloni, C.; Fusco, F.R. Modulation of Phospho-CREB by Systemically Administered Recombinant BDNF in the Hippocampus of the R6/2 Mouse Model of Huntington’s Disease. Neurosci. J. 2019, 2019, 8363274. [Google Scholar] [CrossRef]
  97. Silva, A.; Naia, L.; Dominguez, A.; Ribeiro, M.; Rodrigues, J.; Vieira, O.V.; Lessmann, V.; Rego, A.C. Overexpression of BDNF and Full-Length TrkB Receptor Ameliorate Striatal Neural Survival in Huntington’s Disease. Neurodegener. Dis. 2015, 15, 207–218. [Google Scholar] [CrossRef]
  98. Ma, Q.; Yang, J.M.; Li, T.; Milner, T.A.; Hempstead, B.L. Selective reduction of striatal mature BDNF without induction of proBDNF in the zQ175 mouse model of Huntington’s disease. Neurobiol. Dis. 2015, 82, 466–477. [Google Scholar] [CrossRef] [PubMed]
  99. Plotkin, J.L.; Day, M.; Peterson, J.D.; Xie, Z.; Kress, G.J.; Rafalovich, I.; Kondapalli, J.; Gertler, T.S.; Flajolet, M.; Greengard, P.; et al. Impaired TrkB receptor signaling underlies corticostriatal dysfunction in Huntington’s disease. Neuron 2014, 83, 178–188. [Google Scholar] [CrossRef]
  100. Griffioen, K.J.; Wan, R.; Brown, T.R.; Okun, E.; Camandola, S.; Mughal, M.R.; Phillips, T.M.; Mattson, M.P. Aberrant heart rate and brainstem brain-derived neurotrophic factor (BDNF) signaling in a mouse model of Huntington’s disease. Neurobiol. Aging 2012, 33, 1481.e1–1481.e5. [Google Scholar] [CrossRef]
  101. Saleh, N.; Moutereau, S.; Durr, A.; Krystkowiak, P.; Azulay, J.P.; Tranchant, C.; Broussolle, E.; Morin, F.; Bachoud-Levi, A.C.; Maison, P. Neuroendocrine disturbances in Huntington’s disease. PLoS ONE 2009, 4, e4962. [Google Scholar] [CrossRef]
  102. Smarr, B.; Cutler, T.; Loh, D.H.; Kudo, T.; Kuljis, D.; Kriegsfeld, L.; Ghiani, C.A.; Colwell, C.S. Circadian dysfunction in the Q175 model of Huntington’s disease: Network analysis. J. Neurosci. Res. 2019, 97, 1606–1623. [Google Scholar] [CrossRef] [PubMed]
  103. Liang, F.Q.; Sohrabji, F.; Miranda, R.; Earnest, B.; Earnest, D. Expression of brain-derived neurotrophic factor and its cognate receptor, TrkB, in the rat suprachiasmatic nucleus. Exp. Neurol. 1998, 151, 184–193. [Google Scholar] [CrossRef] [PubMed]
  104. Pan, Y.; Daito, T.; Sasaki, Y.; Chung, Y.H.; Xing, X.; Pondugula, S.; Swamidass, S.J.; Wang, T.; Kim, A.H.; Yano, H. Inhibition of DNA Methyltransferases Blocks Mutant Huntingtin-Induced Neurotoxicity. Sci. Rep. 2016, 6, 31022. [Google Scholar] [CrossRef]
  105. Pan, Y.; Zhu, Y.; Yang, W.; Tycksen, E.; Liu, S.; Palucki, J.; Zhu, L.; Sasaki, Y.; Sharma, M.K.; Kim, A.H.; et al. The role of Twist1 in mutant huntingtin-induced transcriptional alterations and neurotoxicity. J. Biol. Chem. 2018, 293, 11850–11866. [Google Scholar] [CrossRef]
  106. Yu, C.; Li, C.H.; Chen, S.; Yoo, H.; Qin, X.; Park, H. Decreased BDNF Release in Cortical Neurons of a Knock-in Mouse Model of Huntington’s Disease. Sci. Rep. 2018, 8, 16976. [Google Scholar] [CrossRef]
  107. Pineda, J.R.; Pardo, R.; Zala, D.; Yu, H.; Humbert, S.; Saudou, F. Genetic and pharmacological inhibition of calcineurin corrects the BDNF transport defect in Huntington’s disease. Mol. Brain 2009, 2, 33. [Google Scholar] [CrossRef] [PubMed]
  108. Maloney, M.T.; Wang, W.; Bhowmick, S.; Millan, I.; Kapur, M.; Herrera, N.; Frost, E.; Zhang, E.Y.; Song, S.T.; Wang, M.; et al. Failure to Thrive: Impaired BDNF Transport along the Cortical-Striatal Axis in Mouse Q140 Neurons of Huntington’s Disease. Biology 2023, 12, 157. [Google Scholar] [CrossRef]
  109. Kaye, J.; Reisine, T.; Finkbeiner, S. Huntington’s disease iPSC models-using human patient cells to understand the pathology caused by expanded CAG repeats. Fac. Rev. 2022, 11, 16. [Google Scholar] [CrossRef]
  110. Mehta, S.R.; Tom, C.M.; Wang, Y.; Bresee, C.; Rushton, D.; Mathkar, P.P.; Tang, J.; Mattis, V.B. Human Huntington’s Disease iPSC-Derived Cortical Neurons Display Altered Transcriptomics, Morphology, and Maturation. Cell Rep. 2018, 25, 1081–1096. [Google Scholar] [CrossRef]
  111. Mattis, V.B.; Tom, C.; Akimov, S.; Saeedian, J.; Ostergaard, M.E.; Southwell, A.L.; Doty, C.N.; Ornelas, L.; Sahabian, A.; Lenaeus, L.; et al. HD iPSC-derived neural progenitors accumulate in culture and are susceptible to BDNF withdrawal due to glutamate toxicity. Hum. Mol. Genet. 2015, 24, 3257–3271. [Google Scholar] [CrossRef]
  112. Victor, M.B.; Richner, M.; Olsen, H.E.; Lee, S.W.; Monteys, A.M.; Ma, C.; Huh, C.J.; Zhang, B.; Davidson, B.L.; Yang, X.W.; et al. Striatal neurons directly converted from Huntington’s disease patient fibroblasts recapitulate age-associated disease phenotypes. Nat. Neurosci. 2018, 21, 341–352. [Google Scholar] [CrossRef]
  113. Monk, R.; Lee, K.; Jones, K.S.; Connor, B. Directly reprogrammed Huntington’s disease neural precursor cells generate striatal neurons exhibiting aggregates and impaired neuronal maturation. Stem Cells 2021, 39, 1410–1422. [Google Scholar] [CrossRef] [PubMed]
  114. Gines, S.; Paoletti, P.; Alberch, J. Impaired TrkB-mediated ERK1/2 activation in huntington disease knock-in striatal cells involves reduced p52/p46 Shc expression. J. Biol. Chem. 2010, 285, 21537–21548. [Google Scholar] [CrossRef] [PubMed]
  115. Zhao, X.; Chen, X.Q.; Han, E.; Hu, Y.; Paik, P.; Ding, Z.; Overman, J.; Lau, A.L.; Shahmoradian, S.H.; Chiu, W.; et al. TRiC subunits enhance BDNF axonal transport and rescue striatal atrophy in Huntington’s disease. Proc. Natl. Acad. Sci. USA 2016, 113, E5655–E5664. [Google Scholar] [CrossRef] [PubMed]
  116. Zuccato, C.; Valenza, M.; Cattaneo, E. Molecular mechanisms and potential therapeutical targets in Huntington’s disease. Physiol. Rev. 2010, 90, 905–981. [Google Scholar] [CrossRef]
  117. Li, X.J.; Li, S.H.; Sharp, A.H.; Nucifora, F.C., Jr.; Schilling, G.; Lanahan, A.; Worley, P.; Snyder, S.H.; Ross, C.A. A huntingtin-associated protein enriched in brain with implications for pathology. Nature 1995, 378, 398–402. [Google Scholar] [CrossRef]
  118. Liot, G.; Zala, D.; Pla, P.; Mottet, G.; Piel, M.; Saudou, F. Mutant Huntingtin alters retrograde transport of TrkB receptors in striatal dendrites. J. Neurosci. 2013, 33, 6298–6309. [Google Scholar] [CrossRef]
  119. Zala, D.; Colin, E.; Rangone, H.; Liot, G.; Humbert, S.; Saudou, F. Phosphorylation of mutant huntingtin at S421 restores anterograde and retrograde transport in neurons. Hum. Mol. Genet. 2008, 17, 3837–3846. [Google Scholar] [CrossRef]
  120. Mees, I.; Tran, H.; Roberts, A.; Lago, L.; Li, S.; Roberts, B.R.; Hannan, A.J.; Renoir, T. Quantitative Phosphoproteomics Reveals Extensive Protein Phosphorylation Dysregulation in the Cerebral Cortex of Huntington’s Disease Mice Prior to Onset of Symptoms. Mol. Neurobiol. 2022, 59, 2456–2471. [Google Scholar] [CrossRef]
  121. Wang, C.S.; Kavalali, E.T.; Monteggia, L.M. BDNF signaling in context: From synaptic regulation to psychiatric disorders. Cell 2022, 185, 62–76. [Google Scholar] [CrossRef]
  122. Loi, S.M.; Tsoukra, P.; Sun, E.; Chen, Z.; Wibawa, P.; Biase, M.D.; Farrand, S.; Eratne, D.; Kelso, W.; Evans, A.; et al. Survival in Huntington’s disease and other young-onset dementias. Int. J. Geriatr. Psychiatry 2023, 38, e5913. [Google Scholar] [CrossRef] [PubMed]
  123. Giralt, A.; Carreton, O.; Lao-Peregrin, C.; Martin, E.D.; Alberch, J. Conditional BDNF release under pathological conditions improves Huntington’s disease pathology by delaying neuronal dysfunction. Mol. Neurodegener. 2011, 6, 71. [Google Scholar] [CrossRef] [PubMed]
  124. Lynch, G.; Kramar, E.A.; Rex, C.S.; Jia, Y.; Chappas, D.; Gall, C.M.; Simmons, D.A. Brain-derived neurotrophic factor restores synaptic plasticity in a knock-in mouse model of Huntington’s disease. J. Neurosci. 2007, 27, 4424–4434. [Google Scholar] [CrossRef]
  125. Xie, Y.; Hayden, M.R.; Xu, B. BDNF overexpression in the forebrain rescues Huntington’s disease phenotypes in YAC128 mice. J. Neurosci. 2010, 30, 14708–14718. [Google Scholar] [CrossRef]
  126. Simmons, D.A.; Rex, C.S.; Palmer, L.; Pandyarajan, V.; Fedulov, V.; Gall, C.M.; Lynch, G. Up-regulating BDNF with an ampakine rescues synaptic plasticity and memory in Huntington’s disease knockin mice. Proc. Natl. Acad. Sci. USA 2009, 106, 4906–4911. [Google Scholar] [CrossRef]
  127. Torres-Cruz, F.M.; Mendoza, E.; Vivar-Cortes, I.C.; Garcia-Sierra, F.; Hernandez-Echeagaray, E. Do BDNF and NT-4/5 exert synergistic or occlusive effects on corticostriatal transmission in a male mouse model of Huntington’s disease? J. Neurosci. Res. 2019, 97, 1665–1677. [Google Scholar] [CrossRef]
  128. Martire, A.; Pepponi, R.; Domenici, M.R.; Ferrante, A.; Chiodi, V.; Popoli, P. BDNF prevents NMDA-induced toxicity in models of Huntington’s disease: The effects are genotype specific and adenosine A2A receptor is involved. J. Neurochem. 2013, 125, 225–235. [Google Scholar] [CrossRef] [PubMed]
  129. 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]
  130. Park, H. Cortical Axonal Secretion of BDNF in the Striatum Is Disrupted in the Mutant-huntingtin Knock-in Mouse Model of Huntington’s Disease. Exp. Neurobiol. 2018, 27, 217–225. [Google Scholar] [CrossRef]
  131. Spies, J.; Covarrubias-Pinto, A.; Carcamo, C.; Arancibia, Y.; Salazar, F.; Paredes-Martinez, C.; Otth, C.; Castro, M.; Zambrano, A. Modulation of Synaptic Plasticity Genes Associated to DNA Damage in a Model of Huntington’s Disease. Neurochem. Res. 2023, 48, 2093–2103. [Google Scholar] [CrossRef]
  132. Brito, V.; Giralt, A.; Enriquez-Barreto, L.; Puigdellivol, M.; Suelves, N.; Zamora-Moratalla, A.; Ballesteros, J.J.; Martin, E.D.; Dominguez-Iturza, N.; Morales, M.; et al. Neurotrophin receptor p75(NTR) mediates Huntington’s disease-associated synaptic and memory dysfunction. J. Clin. Investig. 2014, 124, 4411–4428. [Google Scholar] [CrossRef] [PubMed]
  133. Bruyere, J.; Abada, Y.S.; Vitet, H.; Fontaine, G.; Deloulme, J.C.; Ces, A.; Denarier, E.; Pernet-Gallay, K.; Andrieux, A.; Humbert, S.; et al. Presynaptic APP levels and synaptic homeostasis are regulated by Akt phosphorylation of huntingtin. eLife 2020, 9, e56371. [Google Scholar] [CrossRef] [PubMed]
  134. Burrus, C.J.; McKinstry, S.U.; Kim, N.; Ozlu, M.I.; Santoki, A.V.; Fang, F.Y.; Ma, A.; Karadeniz, Y.B.; Worthington, A.K.; Dragatsis, I.; et al. Striatal Projection Neurons Require Huntingtin for Synaptic Connectivity and Survival. Cell Rep. 2020, 30, 642–657.e6. [Google Scholar] [CrossRef] [PubMed]
  135. Wennagel, D.; Braz, B.Y.; Capizzi, M.; Barnat, M.; Humbert, S. Huntingtin coordinates dendritic spine morphology and function through cofilin-mediated control of the actin cytoskeleton. Cell Rep. 2022, 40, 111261. [Google Scholar] [CrossRef]
  136. Cozzolino, F.; Vezzoli, E.; Cheroni, C.; Besusso, D.; Conforti, P.; Valenza, M.; Iacobucci, I.; Monaco, V.; Birolini, G.; Bombaci, M.; et al. ADAM10 hyperactivation acts on piccolo to deplete synaptic vesicle stores in Huntington’s disease. Hum. Mol. Genet. 2021, 30, 1175–1187. [Google Scholar] [CrossRef]
  137. Vezzoli, E.; Caron, I.; Talpo, F.; Besusso, D.; Conforti, P.; Battaglia, E.; Sogne, E.; Falqui, A.; Petricca, L.; Verani, M.; et al. Inhibiting pathologically active ADAM10 rescues synaptic and cognitive decline in Huntington’s disease. J. Clin. Investig. 2019, 129, 2390–2403. [Google Scholar] [CrossRef]
  138. Barron, J.C.; Nafar, F.; Parsons, M.P. Huntingtin is essential for synaptic plasticity in the adult hippocampus. bioRxiv 2020. [Google Scholar] [CrossRef]
  139. McAdam, R.L.; Morton, A.; Gordon, S.L.; Alterman, J.F.; Khvorova, A.; Cousin, M.A.; Smillie, K.J. Loss of huntingtin function slows synaptic vesicle endocytosis in striatal neurons from the htt(Q140/Q140) mouse model of Huntington’s disease. Neurobiol. Dis. 2020, 134, 104637. [Google Scholar] [CrossRef]
  140. Dickson, E.; Soylu-Kucharz, R.; Petersen, A.; Bjorkqvist, M. Hypothalamic expression of huntingtin causes distinct metabolic changes in Huntington’s disease mice. Mol. Metab. 2022, 57, 101439. [Google Scholar] [CrossRef]
  141. Zadel, M.; Maver, A.; Kovanda, A.; Peterlin, B. Transcriptomic Biomarkers for Huntington’s Disease: Are Gene Expression Signatures in Whole Blood Reliable Biomarkers? OMICS 2018, 22, 283–294. [Google Scholar] [CrossRef]
  142. Dunah, A.W.; Jeong, H.; Griffin, A.; Kim, Y.M.; Standaert, D.G.; Hersch, S.M.; Mouradian, M.M.; Young, A.B.; Tanese, N.; Krainc, D. Sp1 and TAFII130 transcriptional activity disrupted in early Huntington’s disease. Science 2002, 296, 2238–2243. [Google Scholar] [CrossRef] [PubMed]
  143. Kuhn, A.; Goldstein, D.R.; Hodges, A.; Strand, A.D.; Sengstag, T.; Kooperberg, C.; Becanovic, K.; Pouladi, M.A.; Sathasivam, K.; Cha, J.H.; et al. Mutant huntingtin’s effects on striatal gene expression in mice recapitulate changes observed in human Huntington’s disease brain and do not differ with mutant huntingtin length or wild-type huntingtin dosage. Hum. Mol. Genet. 2007, 16, 1845–1861. [Google Scholar] [CrossRef] [PubMed]
  144. Huang, C.C.; Faber, P.W.; Persichetti, F.; Mittal, V.; Vonsattel, J.P.; MacDonald, M.E.; Gusella, J.F. Amyloid formation by mutant huntingtin: Threshold, progressivity and recruitment of normal polyglutamine proteins. Somat. Cell. Mol. Genet. 1998, 24, 217–233. [Google Scholar] [CrossRef]
  145. Suhr, S.T.; Senut, M.C.; Whitelegge, J.P.; Faull, K.F.; Cuizon, D.B.; Gage, F.H. Identities of sequestered proteins in aggregates from cells with induced polyglutamine expression. J. Cell Biol. 2001, 153, 283–294. [Google Scholar] [CrossRef] [PubMed]
  146. Benn, C.L.; Sun, T.; Sadri-Vakili, G.; McFarland, K.N.; DiRocco, D.P.; Yohrling, G.J.; Clark, T.W.; Bouzou, B.; Cha, J.H. Huntingtin modulates transcription, occupies gene promoters in vivo, and binds directly to DNA in a polyglutamine-dependent manner. J. Neurosci. 2008, 28, 10720–10733. [Google Scholar] [CrossRef]
  147. Pearl, J.; Shetty, A.; Cantle, J.; Bergey, D.; Bragg, R.; Coffey, S.; Kordasiewicz, H.; Hood, L.; Price, N.; Ament, S.; et al. Altered Huntingtin-Chromatin Interactions Predict Transcriptional and Epigenetic Changes in Huntington’s Disease. bioRxiv 2020. [Google Scholar] [CrossRef]
  148. Pruunsild, P.; Kazantseva, A.; Aid, T.; Palm, K.; Timmusk, T. Dissecting the human BDNF locus: Bidirectional transcription, complex splicing, and multiple promoters. Genomics 2007, 90, 397–406. [Google Scholar] [CrossRef]
  149. Pattabiraman, P.P.; Tropea, D.; Chiaruttini, C.; Tongiorgi, E.; Cattaneo, A.; Domenici, L. Neuronal activity regulates the developmental expression and subcellular localization of cortical BDNF mRNA isoforms in vivo. Mol. Cell. Neurosci. 2005, 28, 556–570. [Google Scholar] [CrossRef]
  150. O’Neill, K.M.; Donohue, K.E.; Omelchenko, A.; Firestein, B.L. The 3’ UTRs of Brain-Derived Neurotrophic Factor Transcripts Differentially Regulate the Dendritic Arbor. Front. Cell Neurosci. 2018, 12, 60. [Google Scholar] [CrossRef]
  151. An, J.J.; Gharami, K.; Liao, G.Y.; Woo, N.H.; Lau, A.G.; Vanevski, F.; Torre, E.R.; Jones, K.R.; Feng, Y.; Lu, B.; et al. Distinct role of long 3’ UTR BDNF mRNA in spine morphology and synaptic plasticity in hippocampal neurons. Cell 2008, 134, 175–187. [Google Scholar] [CrossRef]
  152. Zuccato, C.; Cattaneo, E. Brain-derived neurotrophic factor in neurodegenerative diseases. Nat. Rev. Neurol. 2009, 5, 311–322. [Google Scholar] [CrossRef]
  153. Zuccato, C.; Marullo, M.; Conforti, P.; MacDonald, M.E.; Tartari, M.; Cattaneo, E. Systematic assessment of BDNF and its receptor levels in human cortices affected by Huntington’s disease. Brain Pathol. 2008, 18, 225–238. [Google Scholar] [CrossRef]
  154. Schoenherr, C.J.; Anderson, D.J. The neuron-restrictive silencer factor (NRSF): A coordinate repressor of multiple neuron-specific genes. Science 1995, 267, 1360–1363. [Google Scholar] [CrossRef] [PubMed]
  155. Shimojo, M. Huntingtin regulates RE1-silencing transcription factor/neuron-restrictive silencer factor (REST/NRSF) nuclear trafficking indirectly through a complex with REST/NRSF-interacting LIM domain protein (RILP) and dynactin p150 Glued. J. Biol. Chem. 2008, 283, 34880–34886. [Google Scholar] [CrossRef]
  156. Soldati, C.; Bithell, A.; Conforti, P.; Cattaneo, E.; Buckley, N.J. Rescue of gene expression by modified REST decoy oligonucleotides in a cellular model of Huntington’s disease. J. Neurochem. 2011, 116, 415–425. [Google Scholar] [CrossRef]
  157. Finkbeiner, S.; Tavazoie, S.F.; Maloratsky, A.; Jacobs, K.M.; Harris, K.M.; Greenberg, M.E. CREB: A major mediator of neuronal neurotrophin responses. Neuron 1997, 19, 1031–1047. [Google Scholar] [CrossRef]
  158. Nucifora, F.C., Jr.; Sasaki, M.; Peters, M.F.; Huang, H.; Cooper, J.K.; Yamada, M.; Takahashi, H.; Tsuji, S.; Troncoso, J.; Dawson, V.L.; et al. Interference by huntingtin and atrophin-1 with cbp-mediated transcription leading to cellular toxicity. Science 2001, 291, 2423–2428. [Google Scholar] [CrossRef]
  159. Steffan, J.S.; Kazantsev, A.; Spasic-Boskovic, O.; Greenwald, M.; Zhu, Y.Z.; Gohler, H.; Wanker, E.E.; Bates, G.P.; Housman, D.E.; Thompson, L.M. The Huntington’s disease protein interacts with p53 and CREB-binding protein and represses transcription. Proc. Natl. Acad. Sci. USA 2000, 97, 6763–6768. [Google Scholar] [CrossRef] [PubMed]
  160. McFarland, K.N.; Huizenga, M.N.; Darnell, S.B.; Sangrey, G.R.; Berezovska, O.; Cha, J.H.; Outeiro, T.F.; Sadri-Vakili, G. MeCP2: A novel Huntingtin interactor. Hum. Mol. Genet. 2014, 23, 1036–1044. [Google Scholar] [CrossRef] [PubMed]
  161. Baj, G.; Leone, E.; Chao, M.V.; Tongiorgi, E. Spatial segregation of BDNF transcripts enables BDNF to differentially shape distinct dendritic compartments. Proc. Natl. Acad. Sci. USA 2011, 108, 16813–16818. [Google Scholar] [CrossRef]
  162. Patel, R.; Brophy, C.; Hickling, M.; Neve, J.; Furger, A. Alternative cleavage and polyadenylation of genes associated with protein turnover and mitochondrial function are deregulated in Parkinson’s, Alzheimer’s and ALS disease. BMC Med. Genom. 2019, 12, 60. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, Y.; Sun, Z.; Jia, J.; Du, T.; Zhang, N.; Tang, Y.; Fang, Y.; Fang, D. Overview of Histone Modification. Adv. Exp. Med. Biol. 2021, 1283, 1–16. [Google Scholar] [CrossRef] [PubMed]
  164. van Mierlo, G.; Veenstra, G.J.C.; Vermeulen, M.; Marks, H. The Complexity of PRC2 Subcomplexes. Trends Cell Biol. 2019, 29, 660–671. [Google Scholar] [CrossRef]
  165. McFarland, K.N.; Das, S.; Sun, T.T.; Leyfer, D.; Kim, M.O.; Xia, E.; Sangrey, G.R.; Kuhn, A.; Luthi-Carter, R.; Clark, T.W.; et al. Genome-wide increase in histone H2A ubiquitylation in a mouse model of Huntington’s disease. J. Huntington’s Dis. 2013, 2, 263–277. [Google Scholar] [CrossRef]
  166. Biagioli, M.; Ferrari, F.; Mendenhall, E.M.; Zhang, Y.; Erdin, S.; Vijayvargia, R.; Vallabh, S.M.; Solomos, N.; Manavalan, P.; Ragavendran, A.; et al. Htt CAG repeat expansion confers pleiotropic gains of mutant huntingtin function in chromatin regulation. Hum. Mol. Genet. 2015, 24, 2442–2457. [Google Scholar] [CrossRef]
  167. Merienne, N.; Meunier, C.; Schneider, A.; Seguin, J.; Nair, S.S.; Rocher, A.B.; Le Gras, S.; Keime, C.; Faull, R.; Pellerin, L.; et al. Cell-Type-Specific Gene Expression Profiling in Adult Mouse Brain Reveals Normal and Disease-State Signatures. Cell Rep. 2019, 26, 2477–2493.e1–e9. [Google Scholar] [CrossRef] [PubMed]
  168. Yildirim, F.; Ng, C.W.; Kappes, V.; Ehrenberger, T.; Rigby, S.K.; Stivanello, V.; Gipson, T.A.; Soltis, A.R.; Vanhoutte, P.; Caboche, J.; et al. Early epigenomic and transcriptional changes reveal Elk-1 transcription factor as a therapeutic target in Huntington’s disease. Proc. Natl. Acad. Sci. USA 2019, 116, 24840–24851. [Google Scholar] [CrossRef]
  169. Vashishtha, M.; Ng, C.W.; Yildirim, F.; Gipson, T.A.; Kratter, I.H.; Bodai, L.; Song, W.; Lau, A.; Labadorf, A.; Vogel-Ciernia, A.; et al. Targeting H3K4 trimethylation in Huntington disease. Proc. Natl. Acad. Sci. USA 2013, 110, E3027–E3036. [Google Scholar] [CrossRef]
  170. Valor, L.M.; Guiretti, D.; Lopez-Atalaya, J.P.; Barco, A. Genomic landscape of transcriptional and epigenetic dysregulation in early onset polyglutamine disease. J. Neurosci. 2013, 33, 10471–10482. [Google Scholar] [CrossRef] [PubMed]
  171. Hu, E.; Du, H.; Zhu, X.; Wang, L.; Shang, S.; Wu, X.; Lu, H.; Lu, X. Beta-hydroxybutyrate Promotes the Expression of BDNF in Hippocampal Neurons under Adequate Glucose Supply. Neuroscience 2018, 386, 315–325. [Google Scholar] [CrossRef]
  172. Wenzel, E.D.; Speidell, A.; Flowers, S.A.; Wu, C.; Avdoshina, V.; Mocchetti, I. Histone deacetylase 6 inhibition rescues axonal transport impairments and prevents the neurotoxicity of HIV-1 envelope protein gp120. Cell Death Dis. 2019, 10, 674. [Google Scholar] [CrossRef] [PubMed]
  173. Li, E.; Choi, J.; Sim, H.R.; Kim, J.; Jun, J.H.; Kyung, J.; Ha, N.; Kim, S.; Ryu, K.H.; Chung, S.S.; et al. A novel HDAC6 inhibitor, CKD-504, is effective in treating preclinical models of huntington’s disease. BMB Rep. 2023, 56, 178–183. [Google Scholar] [CrossRef] [PubMed]
  174. Hecklau, K.; Mueller, S.; Koch, S.P.; Mehkary, M.H.; Kilic, B.; Harms, C.; Boehm-Sturm, P.; Yildirim, F. The Effects of Selective Inhibition of Histone Deacetylase 1 and 3 in Huntington’s Disease Mice. Front. Mol. Neurosci. 2021, 14, 616886. [Google Scholar] [CrossRef] [PubMed]
  175. Jia, H.; Wang, Y.; Morris, C.D.; Jacques, V.; Gottesfeld, J.M.; Rusche, J.R.; Thomas, E.A. The Effects of Pharmacological Inhibition of Histone Deacetylase 3 (HDAC3) in Huntington’s Disease Mice. PLoS ONE 2016, 11, e0152498. [Google Scholar] [CrossRef] [PubMed]
  176. Suelves, N.; Kirkham-McCarthy, L.; Lahue, R.S.; Gines, S. A selective inhibitor of histone deacetylase 3 prevents cognitive deficits and suppresses striatal CAG repeat expansions in Huntington’s disease mice. Sci. Rep. 2017, 7, 6082. [Google Scholar] [CrossRef] [PubMed]
  177. Jia, H.; Morris, C.D.; Williams, R.M.; Loring, J.F.; Thomas, E.A. HDAC inhibition imparts beneficial transgenerational effects in Huntington’s disease mice via altered DNA and histone methylation. Proc. Natl. Acad. Sci. USA 2015, 112, E56–E64. [Google Scholar] [CrossRef]
  178. Price, A.J.; Collado-Torres, L.; Ivanov, N.A.; Xia, W.; Burke, E.E.; Shin, J.H.; Tao, R.; Ma, L.; Jia, Y.; Hyde, T.M.; et al. Divergent neuronal DNA methylation patterns across human cortical development reveal critical periods and a unique role of CpH methylation. Genome Biol. 2019, 20, 196. [Google Scholar] [CrossRef]
  179. Ng, C.W.; Yildirim, F.; Yap, Y.S.; Dalin, S.; Matthews, B.J.; Velez, P.J.; Labadorf, A.; Housman, D.E.; Fraenkel, E. Extensive changes in DNA methylation are associated with expression of mutant huntingtin. Proc. Natl. Acad. Sci. USA 2013, 110, 2354–2359. [Google Scholar] [CrossRef]
  180. Lu, A.T.; Narayan, P.; Grant, M.J.; Langfelder, P.; Wang, N.; Kwak, S.; Wilkinson, H.; Chen, R.Z.; Chen, J.; Simon Bawden, C.; et al. DNA methylation study of Huntington’s disease and motor progression in patients and in animal models. Nat. Commun. 2020, 11, 4529. [Google Scholar] [CrossRef]
  181. Lieb, K.; Dreimuller, N.; Wagner, S.; Schlicht, K.; Falter, T.; Neyazi, A.; Muller-Engling, L.; Bleich, S.; Tadic, A.; Frieling, H. BDNF Plasma Levels and BDNF Exon IV Promoter Methylation as Predictors for Antidepressant Treatment Response. Front. Psychiatry 2018, 9, 511. [Google Scholar] [CrossRef]
  182. Iamjan, S.A.; Thanoi, S.; Watiktinkorn, P.; Fachim, H.; Dalton, C.F.; Nudmamud-Thanoi, S.; Reynolds, G.P. Changes of BDNF exon IV DNA methylation are associated with methamphetamine dependence. Epigenomics 2021, 13, 953–965. [Google Scholar] [CrossRef]
  183. Kouter, K.; Nikolac Perkovic, M.; Nedic Erjavec, G.; Milos, T.; Tudor, L.; Uzun, S.; Mimica, N.; Pivac, N.; Videtic Paska, A. Difference in Methylation and Expression of Brain-Derived Neurotrophic Factor in Alzheimer’s Disease and Mild Cognitive Impairment. Biomedicines 2023, 11, 235. [Google Scholar] [CrossRef] [PubMed]
  184. Treble-Barna, A.; Heinsberg, L.W.; Stec, Z.; Breazeale, S.; Davis, T.S.; Kesbhat, A.A.; Chattopadhyay, A.; VonVille, H.M.; Ketchum, A.M.; Yeates, K.O.; et al. Brain-derived neurotrophic factor (BDNF) epigenomic modifications and brain-related phenotypes in humans: A systematic review. Neurosci. Biobehav. Rev. 2023, 147, 105078. [Google Scholar] [CrossRef] [PubMed]
  185. Kumar, S.; Chinnusamy, V.; Mohapatra, T. Epigenetics of Modified DNA Bases: 5-Methylcytosine and Beyond. Front. Genet. 2018, 9, 640. [Google Scholar] [CrossRef] [PubMed]
  186. Kriaucionis, S.; Heintz, N. The nuclear DNA base 5-hydroxymethylcytosine is present in Purkinje neurons and the brain. Science 2009, 324, 929–930. [Google Scholar] [CrossRef]
  187. Clemens, A.W.; Gabel, H.W. Emerging Insights into the Distinctive Neuronal Methylome. Trends Genet. 2020, 36, 816–832. [Google Scholar] [CrossRef]
  188. Loh, Y.E.; Koemeter-Cox, A.; Finelli, M.J.; Shen, L.; Friedel, R.H.; Zou, H. Comprehensive mapping of 5-hydroxymethylcytosine epigenetic dynamics in axon regeneration. Epigenetics 2017, 12, 77–92. [Google Scholar] [CrossRef]
  189. Miao, Z.; He, Y.; Xin, N.; Sun, M.; Chen, L.; Lin, L.; Li, J.; Kong, J.; Jin, P.; Xu, X. Altering 5-hydroxymethylcytosine modification impacts ischemic brain injury. Hum. Mol. Genet. 2015, 24, 5855–5866. [Google Scholar] [CrossRef]
  190. Villar-Menendez, I.; Blanch, M.; Tyebji, S.; Pereira-Veiga, T.; Albasanz, J.L.; Martin, M.; Ferrer, I.; Perez-Navarro, E.; Barrachina, M. Increased 5-methylcytosine and decreased 5-hydroxymethylcytosine levels are associated with reduced striatal A2AR levels in Huntington’s disease. Neuromol. Med. 2013, 15, 295–309. [Google Scholar] [CrossRef]
  191. Jeon, S.J.; Rhee, S.Y.; Ryu, J.H.; Cheong, J.H.; Kwon, K.; Yang, S.I.; Park, S.H.; Lee, J.; Kim, H.Y.; Han, S.H.; et al. Activation of adenosine A2A receptor up-regulates BDNF expression in rat primary cortical neurons. Neurochem. Res. 2011, 36, 2259–2269. [Google Scholar] [CrossRef]
  192. Wang, F.; Yang, Y.; Lin, X.; Wang, J.Q.; Wu, Y.S.; Xie, W.; Wang, D.; Zhu, S.; Liao, Y.Q.; Sun, Q.; et al. Genome-wide loss of 5-hmC is a novel epigenetic feature of Huntington’s disease. Hum. Mol. Genet. 2013, 22, 3641–3653. [Google Scholar] [CrossRef] [PubMed]
  193. MacArthur, I.C.; Dawlaty, M.M. TET Enzymes and 5-Hydroxymethylcytosine in Neural Progenitor Cell Biology and Neurodevelopment. Front. Cell Dev. Biol. 2021, 9, 645335. [Google Scholar] [CrossRef] [PubMed]
  194. Barnat, M.; Capizzi, M.; Aparicio, E.; Boluda, S.; Wennagel, D.; Kacher, R.; Kassem, R.; Lenoir, S.; Agasse, F.; Braz, B.Y.; et al. Huntington’s disease alters human neurodevelopment. Science 2020, 369, 787–793. [Google Scholar] [CrossRef] [PubMed]
  195. Cabrera, J.R.; Lucas, J.J. MAP2 Splicing is Altered in Huntington’s Disease. Brain Pathol. 2017, 27, 181–189. [Google Scholar] [CrossRef]
  196. Lin, L.; Park, J.W.; Ramachandran, S.; Zhang, Y.; Tseng, Y.T.; Shen, S.; Waldvogel, H.J.; Curtis, M.A.; Faull, R.L.; Troncoso, J.C.; et al. Transcriptome sequencing reveals aberrant alternative splicing in Huntington’s disease. Hum. Mol. Genet. 2016, 25, 3454–3466. [Google Scholar] [CrossRef]
  197. Fernandez-Nogales, M.; Santos-Galindo, M.; Hernandez, I.H.; Cabrera, J.R.; Lucas, J.J. Faulty splicing and cytoskeleton abnormalities in Huntington’s disease. Brain Pathol. 2016, 26, 772–778. [Google Scholar] [CrossRef]
  198. Neueder, A.; Landles, C.; Ghosh, R.; Howland, D.; Myers, R.H.; Faull, R.L.M.; Tabrizi, S.J.; Bates, G.P. The pathogenic exon 1 HTT protein is produced by incomplete splicing in Huntington’s disease patients. Sci. Rep. 2017, 7, 1307. [Google Scholar] [CrossRef]
  199. Sathasivam, K.; Neueder, A.; Gipson, T.A.; Landles, C.; Benjamin, A.C.; Bondulich, M.K.; Smith, D.L.; Faull, R.L.; Roos, R.A.; Howland, D.; et al. Aberrant splicing of HTT generates the pathogenic exon 1 protein in Huntington disease. Proc. Natl. Acad. Sci. USA 2013, 110, 2366–2370. [Google Scholar] [CrossRef]
  200. Schwab, C.; Arai, T.; Hasegawa, M.; Yu, S.; McGeer, P.L. Colocalization of transactivation-responsive DNA-binding protein 43 and huntingtin in inclusions of Huntington disease. J. Neuropathol. Exp. Neurol. 2008, 67, 1159–1165. [Google Scholar] [CrossRef]
  201. Sanchez, I.I.; Nguyen, T.B.; England, W.E.; Lim, R.G.; Vu, A.Q.; Miramontes, R.; Byrne, L.M.; Markmiller, S.; Lau, A.L.; Orellana, I.; et al. Huntington’s disease mice and human brain tissue exhibit increased G3BP1 granules and TDP43 mislocalization. J. Clin. Investig. 2021, 131, e140723. [Google Scholar] [CrossRef]
  202. Tann, J.Y.; Wong, L.W.; Sajikumar, S.; Ibanez, C.F. Abnormal TDP-43 function impairs activity-dependent BDNF secretion, synaptic plasticity, and cognitive behavior through altered Sortilin splicing. EMBO J. 2019, 38, e100989. [Google Scholar] [CrossRef] [PubMed]
  203. Liu, T.; Im, W.; Mook-Jung, I.; Kim, M. MicroRNA-124 slows down the progression of Huntington’s disease by promoting neurogenesis in the striatum. Neural Regen. Res. 2015, 10, 786–791. [Google Scholar] [CrossRef] [PubMed]
  204. Johnson, R.; Zuccato, C.; Belyaev, N.D.; Guest, D.J.; Cattaneo, E.; Buckley, N.J. A microRNA-based gene dysregulation pathway in Huntington’s disease. Neurobiol. Dis. 2008, 29, 438–445. [Google Scholar] [CrossRef] [PubMed]
  205. Marti, E.; Pantano, L.; Banez-Coronel, M.; Llorens, F.; Minones-Moyano, E.; Porta, S.; Sumoy, L.; Ferrer, I.; Estivill, X. A myriad of miRNA variants in control and Huntington’s disease brain regions detected by massively parallel sequencing. Nucleic Acids Res. 2010, 38, 7219–7235. [Google Scholar] [CrossRef]
  206. Hoss, A.G.; Kartha, V.K.; Dong, X.; Latourelle, J.C.; Dumitriu, A.; Hadzi, T.C.; Macdonald, M.E.; Gusella, J.F.; Akbarian, S.; Chen, J.F.; et al. MicroRNAs located in the Hox gene clusters are implicated in huntington’s disease pathogenesis. PloS Genet. 2014, 10, e1004188. [Google Scholar] [CrossRef]
  207. Muller, S. In silico analysis of regulatory networks underlines the role of miR-10b-5p and its target BDNF in huntington’s disease. Transl. Neurodegener. 2014, 3, 17. [Google Scholar] [CrossRef]
  208. Modarresi, F.; Pedram Fatemi, R.; Razavipour, S.F.; Ricciardi, N.; Makhmutova, M.; Khoury, N.; Magistri, M.; Volmar, C.H.; Wahlestedt, C.; Faghihi, M.A. A novel knockout mouse model of the noncoding antisense Brain-Derived Neurotrophic Factor (Bdnf) gene displays increased endogenous Bdnf protein and improved memory function following exercise. Heliyon 2021, 7, e07570. [Google Scholar] [CrossRef]
  209. Modarresi, F.; Faghihi, M.A.; Lopez-Toledano, M.A.; Fatemi, R.P.; Magistri, M.; Brothers, S.P.; van der Brug, M.P.; Wahlestedt, C. Inhibition of natural antisense transcripts in vivo results in gene-specific transcriptional upregulation. Nat. Biotechnol. 2012, 30, 453–459. [Google Scholar] [CrossRef]
  210. Richard, A.; Frank, S. Deutetrabenazine in the treatment of Huntington’s disease. Neurodegener. Dis. Manag. 2019, 9, 31–37. [Google Scholar] [CrossRef]
  211. Giampa, C.; Montagna, E.; Dato, C.; Melone, M.A.; Bernardi, G.; Fusco, F.R. Systemic delivery of recombinant brain derived neurotrophic factor (BDNF) in the R6/2 mouse model of Huntington’s disease. PLoS ONE 2013, 8, e64037. [Google Scholar] [CrossRef]
  212. da Fonseca, V.S.; da Silva Colla, A.R.; de Paula Nascimento-Castro, C.; Placido, E.; Rosa, J.M.; Farina, M.; Gil-Mohapel, J.; Rodrigues, A.L.S.; Brocardo, P.S. Brain-Derived Neurotrophic Factor Prevents Depressive-like Behaviors in Early-Symptomatic YAC128 Huntington’s Disease Mice. Mol. Neurobiol. 2018, 55, 7201–7215. [Google Scholar] [CrossRef]
  213. Sims, S.K.; Wilken-Resman, B.; Smith, C.J.; Mitchell, A.; McGonegal, L.; Sims-Robinson, C. Brain-Derived Neurotrophic Factor and Nerve Growth Factor Therapeutics for Brain Injury: The Current Translational Challenges in Preclinical and Clinical Research. Neural Plast. 2022, 2022, 3889300. [Google Scholar] [CrossRef]
  214. Miranda-Lourenco, 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] [PubMed]
  215. Cho, S.R.; Benraiss, A.; Chmielnicki, E.; Samdani, A.; Economides, A.; Goldman, S.A. Induction of neostriatal neurogenesis slows disease progression in a transgenic murine model of Huntington disease. J. Clin. Investig. 2007, 117, 2889–2902. [Google Scholar] [CrossRef] [PubMed]
  216. Kells, A.P.; Fong, D.M.; Dragunow, M.; During, M.J.; Young, D.; Connor, B. AAV-mediated gene delivery of BDNF or GDNF is neuroprotective in a model of Huntington disease. Mol. Ther. 2004, 9, 682–688. [Google Scholar] [CrossRef] [PubMed]
  217. Eskandari, N.; Boroujeni, M.E.; Abdollahifar, M.A.; Piryaei, A.; Khodagholi, F.; Mirbehbahani, S.H.; Siroosi, S.; Moghaddam, M.H.; Aliaghaei, A.; Sadeghi, Y. Transplantation of human dental pulp stem cells compensates for striatal atrophy and modulates neuro-inflammation in 3-nitropropionic acid rat model of Huntington’s disease. Neurosci. Res. 2021, 170, 133–144. [Google Scholar] [CrossRef] [PubMed]
  218. Rossignol, J.; Boyer, C.; Leveque, X.; Fink, K.D.; Thinard, R.; Blanchard, F.; Dunbar, G.L.; Lescaudron, L. Mesenchymal stem cell transplantation and DMEM administration in a 3NP rat model of Huntington’s disease: Morphological and behavioral outcomes. Behav. Brain Res. 2011, 217, 369–378. [Google Scholar] [CrossRef]
  219. Wenceslau, C.V.; de Souza, D.M.; Mambelli-Lisboa, N.C.; Ynoue, L.H.; Araldi, R.P.; da Silva, J.M.; Pagani, E.; Haddad, M.S.; Kerkis, I. Restoration of BDNF, DARPP32, and D2R Expression Following Intravenous Infusion of Human Immature Dental Pulp Stem Cells in Huntington’s Disease 3-NP Rat Model. Cells 2022, 11, 1664. [Google Scholar] [CrossRef]
  220. Pollock, K.; Dahlenburg, H.; Nelson, H.; Fink, K.D.; Cary, W.; Hendrix, K.; Annett, G.; Torrest, A.; Deng, P.; Gutierrez, J.; et al. Human Mesenchymal Stem Cells Genetically Engineered to Overexpress Brain-derived Neurotrophic Factor Improve Outcomes in Huntington’s Disease Mouse Models. Mol. Ther. 2016, 24, 965–977. [Google Scholar] [CrossRef]
  221. Reidling, J.C.; Relano-Gines, A.; Holley, S.M.; Ochaba, J.; Moore, C.; Fury, B.; Lau, A.; Tran, A.H.; Yeung, S.; Salamati, D.; et al. Human Neural Stem Cell Transplantation Rescues Functional Deficits in R6/2 and Q140 Huntington’s Disease Mice. Stem Cell Rep. 2018, 10, 58–72. [Google Scholar] [CrossRef]
  222. Yoon, Y.; Kim, H.S.; Jeon, I.; Noh, J.E.; Park, H.J.; Lee, S.; Park, I.H.; Stevanato, L.; Hicks, C.; Corteling, R.; et al. Implantation of the clinical-grade human neural stem cell line, CTX0E03, rescues the behavioral and pathological deficits in the quinolinic acid-lesioned rodent model of Huntington’s disease. Stem Cells 2020, 38, 936–947. [Google Scholar] [CrossRef]
  223. Kim, H.S.; Jeon, I.; Noh, J.E.; Lee, H.; Hong, K.S.; Lee, N.; Pei, Z.; Song, J. Intracerebral Transplantation of BDNF-overexpressing Human Neural Stem Cells (HB1.F3.BDNF) Promotes Migration, Differentiation and Functional Recovery in a Rodent Model of Huntington’s Disease. Exp. Neurobiol. 2020, 29, 130–137. [Google Scholar] [CrossRef]
  224. Cofano, F.; Boido, M.; Monticelli, M.; Zenga, F.; Ducati, A.; Vercelli, A.; Garbossa, D. Mesenchymal Stem Cells for Spinal Cord Injury: Current Options, Limitations, and Future of Cell Therapy. Int. J. Mol. Sci. 2019, 20, 2698. [Google Scholar] [CrossRef]
  225. Zeng, C.W. Multipotent Mesenchymal Stem Cell-Based Therapies for Spinal Cord Injury: Current Progress and Future Prospects. Biology 2023, 12, 653. [Google Scholar] [CrossRef]
  226. Spires, T.L.; Grote, H.E.; Varshney, N.K.; Cordery, P.M.; van Dellen, A.; Blakemore, C.; Hannan, A.J. Environmental enrichment rescues protein deficits in a mouse model of Huntington’s disease, indicating a possible disease mechanism. J. Neurosci. 2004, 24, 2270–2276. [Google Scholar] [CrossRef] [PubMed]
  227. Zajac, M.S.; Pang, T.Y.; Wong, N.; Weinrich, B.; Leang, L.S.; Craig, J.M.; Saffery, R.; Hannan, A.J. Wheel running and environmental enrichment differentially modify exon-specific BDNF expression in the hippocampus of wild-type and pre-motor symptomatic male and female Huntington’s disease mice. Hippocampus 2010, 20, 621–636. [Google Scholar] [CrossRef] [PubMed]
  228. Casella, C.; Bourbon-Teles, J.; Bells, S.; Coulthard, E.; Parker, G.D.; Rosser, A.; Jones, D.K.; Metzler-Baddeley, C. Drumming Motor Sequence Training Induces Apparent Myelin Remodelling in Huntington’s Disease: A Longitudinal Diffusion MRI and Quantitative Magnetization Transfer Study. J. Huntington’s Dis. 2020, 9, 303–320. [Google Scholar] [CrossRef] [PubMed]
  229. van Bruggen-Rufi, M.C.; Vink, A.C.; Wolterbeek, R.; Achterberg, W.P.; Roos, R.A. The Effect of Music Therapy in Patients with Huntington’s Disease: A Randomized Controlled Trial. J. Huntington’s Dis. 2017, 6, 63–72. [Google Scholar] [CrossRef]
  230. Piira, A.; van Walsem, M.R.; Mikalsen, G.; Nilsen, K.H.; Knutsen, S.; Frich, J.C. Effects of a One Year Intensive Multidisciplinary Rehabilitation Program for Patients with Huntington’s Disease: A Prospective Intervention Study. PLoS Curr. 2013, 5. [Google Scholar] [CrossRef]
  231. Bartlett, D.M.; Govus, A.; Rankin, T.; Lampit, A.; Feindel, K.; Poudel, G.; Teo, W.P.; Lo, J.; Georgiou-Karistianis, N.; Ziman, M.R.; et al. The effects of multidisciplinary rehabilitation on neuroimaging, biological, cognitive and motor outcomes in individuals with premanifest Huntington’s disease. J. Neurol. Sci. 2020, 416, 117022. [Google Scholar] [CrossRef] [PubMed]
  232. Wang, H.; Del Mar, N.; Deng, Y.; Reiner, A. Rescue of BDNF expression by the thalamic parafascicular nucleus with chronic treatment with the mGluR2/3 agonist LY379268 may contribute to the LY379268 rescue of enkephalinergic striatal projection neurons in R6/2 Huntington’s disease mice. Neurosci. Lett. 2021, 763, 136180. [Google Scholar] [CrossRef] [PubMed]
  233. Li, S.H.; Colson, T.L.; Abd-Elrahman, K.S.; Ferguson, S.S.G. Metabotropic Glutamate Receptor 2/3 Activation Improves Motor Performance and Reduces Pathology in Heterozygous zQ175 Huntington Disease Mice. J. Pharmacol. Exp. Ther. 2021, 379, 74–84. [Google Scholar] [CrossRef] [PubMed]
  234. El-Shamarka, M.E.; El-Sahar, A.E.; Saad, M.A.; Assaf, N.; Sayed, R.H. Inosine attenuates 3-nitropropionic acid-induced Huntington’s disease-like symptoms in rats via the activation of the A2AR/BDNF/TrKB/ERK/CREB signaling pathway. Life Sci. 2022, 300, 120569. [Google Scholar] [CrossRef] [PubMed]
  235. Di Pardo, A.; Castaldo, S.; Amico, E.; Pepe, G.; Marracino, F.; Capocci, L.; Giovannelli, A.; Madonna, M.; van Bergeijk, J.; Buttari, F.; et al. Stimulation of S1PR5 with A-971432, a selective agonist, preserves blood-brain barrier integrity and exerts therapeutic effect in an animal model of Huntington’s disease. Hum. Mol. Genet. 2018, 27, 2490–2501. [Google Scholar] [CrossRef] [PubMed]
  236. Simmons, D.A.; Belichenko, N.P.; Yang, T.; Condon, C.; Monbureau, M.; Shamloo, M.; Jing, D.; Massa, S.M.; Longo, F.M. A small molecule TrkB ligand reduces motor impairment and neuropathology in R6/2 and BACHD mouse models of Huntington’s disease. J. Neurosci. 2013, 33, 18712–18727. [Google Scholar] [CrossRef]
  237. Jiang, M.; Peng, Q.; Liu, X.; Jin, J.; Hou, Z.; Zhang, J.; Mori, S.; Ross, C.A.; Ye, K.; Duan, W. Small-molecule TrkB receptor agonists improve motor function and extend survival in a mouse model of Huntington’s disease. Hum. Mol. Genet. 2013, 22, 2462–2470. [Google Scholar] [CrossRef] [PubMed]
  238. Simmons, D.A.; Belichenko, N.P.; Ford, E.C.; Semaan, S.; Monbureau, M.; Aiyaswamy, S.; Holman, C.M.; Condon, C.; Shamloo, M.; Massa, S.M.; et al. A small molecule p75NTR ligand normalizes signalling and reduces Huntington’s disease phenotypes in R6/2 and BACHD mice. Hum. Mol. Genet. 2016, 25, 4920–4938. [Google Scholar] [CrossRef]
  239. Conforti, P.; Zuccato, C.; Gaudenzi, G.; Ieraci, A.; Camnasio, S.; Buckley, N.J.; Mutti, C.; Cotelli, F.; Contini, A.; Cattaneo, E. Binding of the repressor complex REST-mSIN3b by small molecules restores neuronal gene transcription in Huntington’s disease models. J. Neurochem. 2013, 127, 22–35. [Google Scholar] [CrossRef]
  240. Kusko, R.; Dreymann, J.; Ross, J.; Cha, Y.; Escalante-Chong, R.; Garcia-Miralles, M.; Tan, L.J.; Burczynski, M.E.; Zeskind, B.; Laifenfeld, D.; et al. Large-scale transcriptomic analysis reveals that pridopidine reverses aberrant gene expression and activates neuroprotective pathways in the YAC128 HD mouse. Mol. Neurodegener. 2018, 13, 25. [Google Scholar] [CrossRef]
  241. Eddings, C.R.; Arbez, N.; Akimov, S.; Geva, M.; Hayden, M.R.; Ross, C.A. Pridopidine protects neurons from mutant-huntingtin toxicity via the sigma-1 receptor. Neurobiol. Dis. 2019, 129, 118–129. [Google Scholar] [CrossRef]
  242. Shenkman, M.; Geva, M.; Gershoni-Emek, N.; Hayden, M.R.; Lederkremer, G.Z. Pridopidine reduces mutant huntingtin-induced endoplasmic reticulum stress by modulation of the Sigma-1 receptor. J. Neurochem. 2021, 158, 467–481. [Google Scholar] [CrossRef]
  243. Asla, M.M.; Nawar, A.A.; Abdelsalam, A.; Elsayed, E.; Rizk, M.A.; Hussein, M.A.; Kamel, W.A. The Efficacy and Safety of Pridopidine on Treatment of Patients with Huntington’s Disease: A Systematic Review and Meta-Analysis. Mov. Disord. Clin. Pract. 2022, 9, 20–30. [Google Scholar] [CrossRef] [PubMed]
  244. Reilmann, R.; McGarry, A.; Grachev, I.D.; Savola, J.M.; Borowsky, B.; Eyal, E.; Gross, N.; Langbehn, D.; Schubert, R.; Wickenberg, A.T.; et al. Safety and efficacy of pridopidine in patients with Huntington’s disease (PRIDE-HD): A phase 2, randomised, placebo-controlled, multicentre, dose-ranging study. Lancet Neurol. 2019, 18, 165–176. [Google Scholar] [CrossRef] [PubMed]
  245. Couly, S.; Paucard, A.; Bonneaud, N.; Maurice, T.; Benigno, L.; Jourdan, C.; Cohen-Solal, C.; Vignes, M.; Maschat, F. Improvement of BDNF signalling by P42 peptide in Huntington’s disease. Hum. Mol. Genet. 2018, 27, 3012–3028. [Google Scholar] [CrossRef] [PubMed]
  246. Couly, S.; Carles, A.; Denus, M.; Benigno-Anton, L.; Maschat, F.; Maurice, T. Exposure of R6/2 mice in an enriched environment augments P42 therapy efficacy on Huntington’s disease progression. Neuropharmacology 2021, 186, 108467. [Google Scholar] [CrossRef] [PubMed]
  247. Soles-Tarres, I.; Cabezas-Llobet, N.; Lefranc, B.; Leprince, J.; Alberch, J.; Vaudry, D.; Xifro, X. Pituitary Adenylate Cyclase-Activating Polypeptide (PACAP) Protects Striatal Cells and Improves Motor Function in Huntington’s Disease Models: Role of PAC1 Receptor. Front. Pharmacol. 2021, 12, 797541. [Google Scholar] [CrossRef] [PubMed]
  248. Mehta, R.I.; Carpenter, J.S.; Mehta, R.I.; Haut, M.W.; Ranjan, M.; Najib, U.; Lockman, P.; Wang, P.; D’Haese, P.F.; Rezai, A.R. Blood-Brain Barrier Opening with MRI-guided Focused Ultrasound Elicits Meningeal Venous Permeability in Humans with Early Alzheimer Disease. Radiology 2021, 298, 654–662. [Google Scholar] [CrossRef]
  249. Chen, H.; Yang, G.Z.; Getachew, H.; Acosta, C.; Sierra Sanchez, C.; Konofagou, E.E. Focused ultrasound-enhanced intranasal brain delivery of brain-derived neurotrophic factor. Sci. Rep. 2016, 6, 28599. [Google Scholar] [CrossRef]
  250. Hogan, R.E.; Tarquinio, D.; Sperling, M.R.; Klein, P.; Miller, I.; Segal, E.B.; Rabinowicz, A.L.; Carrazana, E. Pharmacokinetics and safety of VALTOCO (NRL-1; diazepam nasal spray) in patients with epilepsy during seizure (ictal/peri-ictal) and nonseizure (interictal) conditions: A phase 1, open-label study. Epilepsia 2020, 61, 935–943. [Google Scholar] [CrossRef]
  251. Wheless, J.W.; Miller, I.; Hogan, R.E.; Dlugos, D.; Biton, V.; Cascino, G.D.; Sperling, M.R.; Liow, K.; Vazquez, B.; Segal, E.B.; et al. Final results from a Phase 3, long-term, open-label, repeat-dose safety study of diazepam nasal spray for seizure clusters in patients with epilepsy. Epilepsia 2021, 62, 2485–2495. [Google Scholar] [CrossRef] [PubMed]
  252. Kasenda, B.; Konig, D.; Manni, M.; Ritschard, R.; Duthaler, U.; Bartoszek, E.; Barenwaldt, A.; Deuster, S.; Hutter, G.; Cordier, D.; et al. Targeting immunoliposomes to EGFR-positive glioblastoma. ESMO Open 2022, 7, 100365. [Google Scholar] [CrossRef] [PubMed]
  253. Saudou, F.; Finkbeiner, S.; Devys, D.; Greenberg, M.E. Huntingtin acts in the nucleus to induce apoptosis but death does not correlate with the formation of intranuclear inclusions. Cell 1998, 95, 55–66. [Google Scholar] [CrossRef] [PubMed]
  254. Fodelianaki, G.; Lansing, F.; Bhattarai, P.; Troullinaki, M.; Zeballos, M.A.; Charalampopoulos, I.; Gravanis, A.; Mirtschink, P.; Chavakis, T.; Alexaki, V.I. Nerve Growth Factor modulates LPS—Induced microglial glycolysis and inflammatory responses. Exp. Cell Res. 2019, 377, 10–16. [Google Scholar] [CrossRef] [PubMed]
  255. Jiang, Z.; Wang, J.; Sun, G.; Feng, M. BDNF-modified human umbilical cord mesenchymal stem cells-derived dopaminergic-like neurons improve rotation behavior of Parkinson’s disease rats through neuroprotection and anti-neuroinflammation. Mol. Cell. Neurosci. 2022, 123, 103784. [Google Scholar] [CrossRef] [PubMed]
  256. Han, R.; Liu, Z.; Sun, N.; Liu, S.; Li, L.; Shen, Y.; Xiu, J.; Xu, Q. BDNF Alleviates Neuroinflammation in the Hippocampus of Type 1 Diabetic Mice via Blocking the Aberrant HMGB1/RAGE/NF-kappaB Pathway. Aging Dis. 2019, 10, 611–625. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Selected mechanisms for BDNF gene dysregulation in HD. BDNF gene regulation is altered in HD and contributes to neuronal death through several mechanisms: (A) Wild-type HTT protein binds with REST in the cytoplasm, which hinders REST nuclear translocation and results in increased neuronal survival, synaptic plasticity, and neurite growth [155]. The reduced affinity of REST to mHTT allows the formation of a repressor complex on cis-regulatory elements of the BDNF exon II transcription start site. CoREST, REST corepressor 1; HDAC, Histone deacetylase; NRSE, Neuron-restrictive silencing element; RE1, Repressor element 1; REST, RE1-silencing transcription factor; SIN3A, SIN3 transcription regulator family member A. (B) The epigenetic dysregulation of BDNF gene transcription in HD neurons due to differential methylation patterns between normal and HD neurons. DNA methyltransferases (DNMT3A and DNMT1) methylate the regulatory elements of the Bdnf exon IV, resulting in the downregulation of its transcription in mHTT-expressing cortical neurons [104,105]. In HD neurons, gene repressive complexes (not shown) may be recruited to methylated cytosine through the action of mHTT and methylcytosine binding protein 2 (MeCP2) to silence Bdnf gene transcription [160].
Figure 1. Selected mechanisms for BDNF gene dysregulation in HD. BDNF gene regulation is altered in HD and contributes to neuronal death through several mechanisms: (A) Wild-type HTT protein binds with REST in the cytoplasm, which hinders REST nuclear translocation and results in increased neuronal survival, synaptic plasticity, and neurite growth [155]. The reduced affinity of REST to mHTT allows the formation of a repressor complex on cis-regulatory elements of the BDNF exon II transcription start site. CoREST, REST corepressor 1; HDAC, Histone deacetylase; NRSE, Neuron-restrictive silencing element; RE1, Repressor element 1; REST, RE1-silencing transcription factor; SIN3A, SIN3 transcription regulator family member A. (B) The epigenetic dysregulation of BDNF gene transcription in HD neurons due to differential methylation patterns between normal and HD neurons. DNA methyltransferases (DNMT3A and DNMT1) methylate the regulatory elements of the Bdnf exon IV, resulting in the downregulation of its transcription in mHTT-expressing cortical neurons [104,105]. In HD neurons, gene repressive complexes (not shown) may be recruited to methylated cytosine through the action of mHTT and methylcytosine binding protein 2 (MeCP2) to silence Bdnf gene transcription [160].
Biomedicines 11 02275 g001
Figure 2. Overview of selected BDNF-directed molecular therapies in HD. Current therapeutic strategies for increasing BDNF expression in HD include (A) AAV-BDNF transduction of astrocytes or engraftment of BDNF-expressing MSCs; (B) small molecules targeting cell-surface G protein-coupled receptors (GPCRs) or ionotropic glutamate receptors; or (C) novel peptidergic compounds or cell-permeable small molecules that increase PKA signaling and CREB phosphorylation. Other approaches include (D) the modulation of ER calcium release or (E) the inhibition of DNMTs. Abbreviations used in alphabetical order: A2AR, Adenosine-2A receptor; AAV, Adeno-associated virus; AMP, Adenosine monophosphate; AMPAR, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; ATP, Adenosine triphosphate; αs, G-protein stimulatory alpha subunit; cAMP, Cyclic adenosine monophosphate; CREB, cAMP response element-binding protein; CBP, CREB-binding protein; DNMT, DNA cytosine-5-methyltransferases; MSCs, Mesenchymal stem cells; Me, Methyl group; mGlu2/3, Metabotropic glutamate receptor 2/3; PDE4, Phosphodiesterase 4b; PKA, protein kinase A; S1PR5, Sphingosine-1-phosphate receptor 5; VIP, Vasoactive intestinal peptide; VIPR1, Vasoactive intestinal peptide receptor 1.
Figure 2. Overview of selected BDNF-directed molecular therapies in HD. Current therapeutic strategies for increasing BDNF expression in HD include (A) AAV-BDNF transduction of astrocytes or engraftment of BDNF-expressing MSCs; (B) small molecules targeting cell-surface G protein-coupled receptors (GPCRs) or ionotropic glutamate receptors; or (C) novel peptidergic compounds or cell-permeable small molecules that increase PKA signaling and CREB phosphorylation. Other approaches include (D) the modulation of ER calcium release or (E) the inhibition of DNMTs. Abbreviations used in alphabetical order: A2AR, Adenosine-2A receptor; AAV, Adeno-associated virus; AMP, Adenosine monophosphate; AMPAR, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; ATP, Adenosine triphosphate; αs, G-protein stimulatory alpha subunit; cAMP, Cyclic adenosine monophosphate; CREB, cAMP response element-binding protein; CBP, CREB-binding protein; DNMT, DNA cytosine-5-methyltransferases; MSCs, Mesenchymal stem cells; Me, Methyl group; mGlu2/3, Metabotropic glutamate receptor 2/3; PDE4, Phosphodiesterase 4b; PKA, protein kinase A; S1PR5, Sphingosine-1-phosphate receptor 5; VIP, Vasoactive intestinal peptide; VIPR1, Vasoactive intestinal peptide receptor 1.
Biomedicines 11 02275 g002
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

Speidell, A.; Bin Abid, N.; Yano, H. Brain-Derived Neurotrophic Factor Dysregulation as an Essential Pathological Feature in Huntington’s Disease: Mechanisms and Potential Therapeutics. Biomedicines 2023, 11, 2275. https://doi.org/10.3390/biomedicines11082275

AMA Style

Speidell A, Bin Abid N, Yano H. Brain-Derived Neurotrophic Factor Dysregulation as an Essential Pathological Feature in Huntington’s Disease: Mechanisms and Potential Therapeutics. Biomedicines. 2023; 11(8):2275. https://doi.org/10.3390/biomedicines11082275

Chicago/Turabian Style

Speidell, Andrew, Noman Bin Abid, and Hiroko Yano. 2023. "Brain-Derived Neurotrophic Factor Dysregulation as an Essential Pathological Feature in Huntington’s Disease: Mechanisms and Potential Therapeutics" Biomedicines 11, no. 8: 2275. https://doi.org/10.3390/biomedicines11082275

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