Next Article in Journal
Vitamin A, D, E, and K as Matrix Metalloproteinase-2/9 Regulators That Affect Expression and Enzymatic Activity
Next Article in Special Issue
Parkinson’s Disease: Cells Succumbing to Lifelong Dopamine-Related Oxidative Stress and Other Bioenergetic Challenges
Previous Article in Journal
Overexpression of an ART1-Interacting Gene OsNAC016 Improves Al Tolerance in Rice
Previous Article in Special Issue
Involvement of Mitochondria in Parkinson’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pathogenic Impact of Fatty Acid-Binding Proteins in Parkinson’s Disease—Potential Biomarkers and Therapeutic Targets

by
Ichiro Kawahata
1,* and
Kohji Fukunaga
1,2
1
Department of CNS Drug Innovation, Graduate School of Pharmaceutical Sciences, Tohoku University, Sendai 980-8578, Japan
2
BRI Pharma Inc., Sendai 982-0804, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(23), 17037; https://doi.org/10.3390/ijms242317037
Submission received: 30 October 2023 / Revised: 26 November 2023 / Accepted: 30 November 2023 / Published: 1 December 2023
(This article belongs to the Special Issue Recent Molecular Research of Parkinson's Disease)

Abstract

:
Parkinson’s disease is a neurodegenerative condition characterized by motor dysfunction resulting from the degeneration of dopamine-producing neurons in the midbrain. This dopamine deficiency gives rise to a spectrum of movement-related symptoms, including tremors, rigidity, and bradykinesia. While the precise etiology of Parkinson’s disease remains elusive, genetic mutations, protein aggregation, inflammatory processes, and oxidative stress are believed to contribute to its development. In this context, fatty acid-binding proteins (FABPs) in the central nervous system, FABP3, FABP5, and FABP7, impact α-synuclein aggregation, neurotoxicity, and neuroinflammation. These FABPs accumulate in mitochondria during neurodegeneration, disrupting their membrane potential and homeostasis. In particular, FABP3, abundant in nigrostriatal dopaminergic neurons, is responsible for α-synuclein propagation into neurons and intracellular accumulation, affecting the loss of mesencephalic tyrosine hydroxylase protein, a rate-limiting enzyme of dopamine biosynthesis. This review summarizes the characteristics of FABP family proteins and delves into the pathogenic significance of FABPs in the pathogenesis of Parkinson’s disease. Furthermore, it examines potential novel therapeutic targets and early diagnostic biomarkers for Parkinson’s disease and related neurodegenerative disorders.

1. Introduction

Parkinson’s disease is characterized by the progressive degeneration of dopaminergic neurons within the midbrain, particularly the nigrostriatal system, which plays a crucial role in motor function regulation. This degenerative process primarily occurs in the substantia nigra, leading to a substantial reduction in dopamine levels that significantly affects motor control, resulting in symptoms like tremors, muscle rigidity, and bradykinesia.
Despite extensive research efforts, the detailed etiology of Parkinson’s disease remains incompletely understood. The disease’s onset involves a complex interplay of genetic and environmental factors, including the abnormal accumulation and aggregation of α-synuclein, the pathogenic protein linked to Parkinson’s disease, as well as inflammatory responses and oxidative stress. While various studies have highlighted the impact of genetic mutations in familial Parkinson’s disease, the mechanisms triggering the disease in sporadic cases without clear genetic influences remain challenging to decipher.
Early therapeutic intervention is crucial for managing Parkinson’s disease; however, predicting its onset in the early stages remains challenging. Previous research indicates the potential utility of fatty acid-binding proteins (FABPs) as diagnostic biomarkers for various conditions, including cerebral infarction, Alzheimer’s disease, dementia with Lewy bodies, and Parkinson’s disease [1,2,3,4,5]. Thus, there remains a need for a deeper understanding of the molecular mechanisms underlying the involvement of FABPs in disease pathogenesis.
Therefore, this review aims to elucidate the mechanisms underlying the selective dysfunction of dopaminergic neurons and consequent neurodegeneration in Parkinson’s disease, focusing on the involvement of FABPs. Specifically, we highlight the emerging role of FABPs in α-synucleinopathies, presenting recent research elucidating novel mechanisms associated with α-synuclein propagation and toxicity through protein aggregation. Additionally, we delve into newfound insights into the molecular pathways governing mitochondrial dysfunction during the progression of neurodegeneration. Drawing from these research findings, we aim to explore the potential of novel pharmaceutical targets and biomarkers for effective disease prediction and therapeutic interventions.

2. Physiological Function of FABP and Involvement in Neurodegenerative Diseases

Fatty acids perform a variety of physiological functions in the body, serving as a source of energy for internal combustion, a major component of cell membranes, and a regulator of inflammatory responses. Excessive intake of fatty acids contributes to energy overload, obesity, and risk of brain inflammation. Fatty acids include saturated fatty acids; unsaturated fatty acids, which include monounsaturated and polyunsaturated fatty acids; and trans fatty acids. Some of these polyunsaturated fatty acids cannot be synthesized in the animal’s body and must be obtained from the diet. An excess of or deficiency in polyunsaturated fatty acids are associated with various functional disorders, neurological symptoms, and an increased risk of disease pathogenesis. Polyunsaturated fatty acids can be classified into omega-3 fatty acids, represented by docosahexaenoic acid (DHA) and eicosapentaenoic acid (EPA), and omega-6 fatty acids, which are metabolized into linoleic acid and arachidonic acid.
Because fatty acids are hydrophobic, they require carrier proteins to transport them to intracellular organelles, and FABPs are responsible for this physiological function [6,7,8,9]. FABP is a protein consisting of approximately 130 amino acid residues, and nine isoforms have been identified in humans [10,11]. Each of these isoforms is known to have some specificity in its expression distribution (Table 1). Three of these isoforms, FABP3, FABP5, and FABP7, are expressed in the nervous system [12,13,14,15]. FABP3, which was first identified in the heart [16,17], is abundantly expressed in mature neurons from the postnatal to the adult stage [14]. In contrast, FABP5 and FABP7 are maximally expressed in the fetus and neonate, during which time they are expressed in glial and neural stem cells [14,18,19]. FABP3 has a high affinity for n-6 polyunsaturated fatty acids such as arachidonic acid, FABP5 has a high affinity for saturated fatty acids such as stearic acid and palmitic acid, and FABP7 has a high affinity for n-3 polyunsaturated fatty acids such as DHA [20,21,22].
Polyunsaturated fatty acid itself affects the toxic expression of α-synuclein, the pathogenic protein of Parkinson’s disease. Previous reports indicate that polyunsaturated fatty acids bind to α-synuclein and promote oligomer formation [23,24,25]. The exposure of cultured mesencephalic neurons to polyunsaturated fatty acids increased α-synuclein oligomer levels [25]. In addition, when mice expressing A53T, a representative mutant family line of α-synuclein in familial Parkinson’s disease, were fed a DHA-containing diet, low concentrations of α-synuclein suppressed neuronal accumulation and toxic expression. In contrast, high concentrations of DHA increased intracellular accumulation of soluble and insoluble α-synuclein and neuronal injury [26]. In humans, fatty acids have also been implicated in the pathogenesis of Lewy body disease, and α-synuclein oligomers in the lipid fraction are detected in autopsy brains of Parkinson’s disease and Lewy body dementia patients but not in healthy subjects [25]. Furthermore, a higher intake of arachidonic acid, an omega-6 polyunsaturated fatty acid, has been suggested to increase the risk of Parkinson’s disease pathogenesis [27].
Furthermore, FABPs have pathogenic impacts and the potential for predictive biomarkers on various diseases, including brain injury and neurodegenerative disorders. FABP3 and FABP7 exhibit distinct distribution patterns within brain tissues, with FABP3 displaying notably higher concentrations in brain injury [28]. Elevated serum levels of FABP7 were observed in patients with Alzheimer’s disease, Parkinson’s disease, and other cognitive disorders [4]. In addition, FABP3 expression in the substantia nigra is known to be increased in autopsy brains of Parkinson’s disease patients [29]. In this context, FABP3 co-localizes with phosphorylated α-synuclein in Lewy bodies [30]. Serum FABP3 levels are increased in patients with Parkinson’s disease and dementia with Lewy bodies compared to healthy controls [3,31]. In addition, serum FABP3 was elevated in dementia with Lewy bodies and Parkinson’s disease with dementia, compared to non-dementia controls [2,32]. These clinical findings suggest that FABP3 may be involved in the pathogenesis of Lewy body diseases, including Parkinson’s disease. Based on these insights, the following chapters will describe the characteristics of Parkinson’s disease, the selective degeneration of dopaminergic neurons, and the pathogenic impact of FABPs in the disease.
Table 1. Tissue distribution and expressed cell types of the fatty acid-binding protein (FABP) subfamilies.
Table 1. Tissue distribution and expressed cell types of the fatty acid-binding protein (FABP) subfamilies.
FABP SubfamilyTissue DistributionExpressed CellsRef
FABP1
(Liver FABP)
Liver, intestine, kidney, pancreasHepatocytes, enterocytes[33,34,35]
FABP2
(Intestinal FABP)
IntestineEnterocytes[36,37,38]
FABP3
(Heart FABP)
Heart, skeletal muscle, brainCardiomyocytes, myocytes, neurons[18,39,40,41,42,43]
FABP4
(Adipocyte FABP)
Adipose tissue, macrophagesAdipocytes, macrophages[44,45,46,47]
FABP5
(Epidermal FABP)
Epidermis, brain, adipose tissueKeratinocytes, adipocytes, glial cells, neurons[18,43,48,49,50]
FABP6
(Ileal FABP)
IntestineEnterocytes[51,52,53,54]
FABP7
(Brain FABP)
Brain, eye, kidney, mammary glandNeural stem cells, oligodendrocytes, astrocytes, ependymal cells[18,41,43,55,56]
FABP8
(Myelin FABP)
Myelin-forming cells in the peripheral nervous systemSchwann cells, oligodendrocytes[43,56]
FABP9
(Testis FABP)
TestisSalivary gland, mammary gland[57,58]

3. Pathology of Parkinson’s Disease and Current Issues

There are over 10 million Parkinson’s disease patients worldwide, accounting for 1–3% of the global population aged 60 and above [59]. Half of those aged 85 and above develop Parkinson’s disease [60]. The evolution of symptomatic therapies for Parkinson’s disease, including dopamine agonists as well as L-DOPA, has been remarkable [61]. However, a fundamental treatment for Parkinson’s disease has yet to be developed. Initiating treatment after onset does not lead to a favorable prognosis. At the onset of Parkinson’s disease, the pathogenic proteins have already accumulated in the brain, and this pathology of dopaminergic neuronal loss induces clinical symptoms. Therefore, there is an expectation for the establishment of methods to predict the onset early.
Parkinson’s disease was first diagnosed by James Parkinson in 1817, detailing its clinical features as tremor, rigidity, bradykinesia, gait disturbances, and postural instability [62]. Cognitive symptoms commonly emerge in the advanced stages of the disease [63]. Approximately 20–40% of all Parkinson’s cases develop Parkinson’s disease dementia, with an average progression time of 10 years. In total, 40% of Parkinson’s patients with severe olfactory dysfunction progress to dementia [64,65]. In this regard, the importance of the olfactory bulb as an entry site for prion-like transmission in neurodegenerative diseases is suggested [66]. Pathologically, Parkinson’s disease is characterized by the loss of dopamine-biosynthesizing neurons in the substantia nigra pars compacta. It is also denoted by the abnormal deposition of the pathogenic protein α-synuclein in the cell body and neuronal processes, forming Lewy bodies and Lewy neurites, respectively [67,68,69]. Parkinson’s disease, Parkinson’s disease dementia, and dementia with Lewy bodies share the accumulation and aggregation of the pathogenic protein α-synuclein within neurons, leading to the formation of Lewy bodies. Therefore, these Lewy body diseases are believed to share common pathological mechanisms.
The decline in motor function observed in Parkinson’s disease arises due to impaired nigrostriatal dopaminergic function [70]. Nigrostriatal dopaminergic projections centrally regulate voluntary movements, and their degeneration contributes to Parkinsonian clinical symptoms. In addition, the dopaminergic system, originating in the substantia nigra pars compacta and the ventral tegmental area, predominantly projecting to the striatum and prefrontal context, significantly influences behavioral activities [71,72,73]. Consequently, lesions in nigral neurons cause concurrent dysfunction of agonist and antagonist muscle pairs in animal models of Parkinsonism [74] and sporadic Parkinson’s disease [75]. The dopaminergic function is regulated by the neurotransmitter dopamine. This catecholamine is biosynthesized from L-tyrosine by the rate-limiting enzyme tyrosine hydroxylase (TH) and aromatic L-amino acid decarboxylase (AADC) [76]. TH requires tetrahydrobiopterin, which is biosynthesized by GTP cyclohydrolase I (GTPCH1), to perform its enzymatic activity [77,78]. Because the enzymatic activity of TH protein strictly controls the rate-limiting step of dopamine biosynthesis [79], unlike those of other dopamine biosynthesizing enzymes, the expression level and activity of TH, which is precisely regulated via phosphorylation, directly affect intracellular dopamine amount.

4. Biochemistry and Pathology of Tyrosine Hydroxylase

In the context of movement disorders, dopamine is an essential neurotransmitter for motor function [80]. In this context, TH is a rate-limiting enzyme for dopamine biosynthesis [79] and is selectively expressed in catecholaminergic neurons in the central nervous system. AADC is responsible for the subsequent reaction to TH [81,82], and GTPCH1 is essential for the biosynthesis of cofactors for TH [83,84], but their enzymatic activities are not strictly regulated. Therefore, TH activity is almost directly related to the amount of DA. Shortly, TH is a homotetramer consisting of four subunits through its C-terminal domain [85]. Its activity is primarily regulated via the phosphorylation of its serine 40 amino acid residue in the N-terminal region [86,87]. Phosphorylation of TH is facilitated mainly by the cAMP-dependent protein kinase (PKA), Ca2+/calmodulin-dependent protein kinase II (CaMKII), and also stress-responsive enzymes, including mitogen-activated protein kinase activated protein kinase (MAPKAPK) and mitogen- and stress-activated kinase 1 (MSK1) [87,88]. Phosphorylated TH is dephosphorylated by a protein phosphatase, such as protein phosphatase 2A (PP2A) [89].
Dopamine-bound TH is inactive and stable, whereas TH is unstable in its phosphorylated active state [88,90,91]. The TH protein exhibits an aggregative nature upon phosphorylation [92,93]. In Parkinson’s disease, the levels of proteasome constituent proteins decrease, leading to reduced activity in the substantia nigra [94,95]. Furthermore, the oligomerization of α-synuclein, a pathogenic protein in Parkinson’s disease, impairs 26S proteasome-mediated protein degradation [96]. In this context, in Parkinson’s disease model cells treated with a proteasome inhibitor, the TH protein forms insoluble aggregates [97]. Specifically, phosphorylated TH at the serine 40 residue demonstrates an aggregation propensity but not dephosphorylated TH (Figure 1A). Indeed, immunohistochemistry reveals that Lewy bodies are immunopositive to TH phosphorylated at the serine 40 in the autopsied brains of Parkinson’s disease patients [98].
On the other hand, TH can be ubiquitinated [100], and the aggregates of phosphorylated TH are ubiquitin-positive [97]. In Parkinson’s disease and dopa-responsive dystonia, the TH protein in the substantia nigra striatum is lost. In model cells with reduced dopamine levels in Parkinson’s disease models or reduced biopterin levels in dopa-responsive dystonia, the phosphorylation level of TH increases, and over time, the TH level decreases [99]. This reduction in TH levels can be inhibited by a proteasome inhibitor. These findings suggest that the phosphorylation of TH is promoted by the decrease in dopamine or biopterin levels, and phosphorylated TH is likely degraded by the proteasome, explaining the mechanism of TH loss in Parkinson’s disease (Figure 1B).

5. Impact of FABP on the Loss of TH Protein and α-Synuclein Aggregation

As mentioned in the previous section, the activation of the TH protein by its phosphorylation accelerates its degradation in Parkinson’s disease model cells as well as the dopa-responsive dystonia model. In this context, the intracellular accumulation of α-synuclein promotes aggregation and phosphorylation, enhancing TH phosphorylation [102,103]. Intriguingly, the cellular uptake and accumulation of α-synuclein in dopaminergic neurons rely on FABP3 [104,105]. FABP3 is abundantly expressed in dopaminergic neurons within the central nervous system [18,42,43,104] and is immunopositive in intracellular inclusions in model cells [104] and in Lewy bodies in patients with Parkinson’s disease [30]. Indeed, in dopaminergic neurons in which FABP3 is knocked out, there is no increase in phosphorylated TH, and loss of TH is suppressed [101]. These data suggest that FABP3 affects the degradation and loss of TH protein through its involvement in α-synuclein aggregate formation and phosphorylation.
In the absence of FABP3, α-synuclein is not taken up by neuronal cells, thus preventing the formation of aggregates [104,105]. Additionally, FABP3 is not only involved in the intracellular uptake and aggregate formation of α-synuclein but also contributes to its propagation within the brain [106,107]. Furthermore, FABP3 is essential for the uptake of both monomeric and fibrillar forms of α-synuclein [105]. Its physiological function requires the presence of dopamine D2 receptors. There are two isoforms of dopamine D2 receptors: the long isoform of the dopamine D2 receptor (D2L receptor) and the short isoform (D2S receptor) [108]. The D2L receptor has 29 more amino acids than D2S and is located in the third intracellular loop. The D2L receptor contains a binding site for FABP3, whereas the D2S receptor lacks this binding site, indicating a selective interaction with FABP3 [109]. Analysis using D2L receptor-specific knockout models suggests that the presence of both FABP3 and D2L receptors is necessary for the uptake of α-synuclein [105]. Additionally, the abundance of D2L receptors in the caveolae structures on the cell membrane is well documented [110,111], and the disruption of these structures leads to a decrease in the physiological function of FABP3 [105]. Based on this evidence, it is hypothesized that the D2L receptor acts as a scaffold protein for FABP3, localizing it near the cell membrane and accelerating the intracellular accumulation of α-synuclein (Figure 2). The expression sites of FABP3 and D2L receptors in the brain, such as the midbrain, olfactory bulb, and cerebral cortex, correspond to the regions where Lewy bodies are observed in Parkinson’s disease and Lewy body dementia [112,113,114]. This correspondence may provide an explanation for the site-selective formation of Lewy bodies in Lewy body diseases.

6. Pathogenic Impact of Fatty Acid-Binding Proteins on Mitochondrial Homeostasis

Mitochondrial dysfunction is critical for the pathogenesis of both familial and sporadic Parkinson’s disease [116]. 1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), a neurotoxin that originated as a byproduct of 1-methyl-4-phenyl-propionoxy-piperidine (MPPP) synthesis, is widely utilized in producing Parkinson’s disease models due to its induction of typical Parkinsonian symptoms effectively managed by L-DOPA [117]. Its ingestion results in pathological findings such as degeneration of substantia nigra dopaminergic neurons and the presence of Lewy bodies, as observed in postmortem brain analyses [118,119]. When MPTP enters the brain, it is oxidized by the enzyme monoamine oxidase B (MAO-B) within astrocytes and microglia to form 1-methyl-4-phenylpyridinium (MPP+), which is then released into the extracellular space and subsequently taken up by dopaminergic neurons. Once inside, MPP+ is internalized into the mitochondria, where it strongly inhibits complex I of the electron transport chain, leading to cellular degeneration due to decreased energy production capacity [120]. Rotenone [121], another well-known mitochondrial toxin, is also widely used to reproduce the Parkinson’s disease phenotype [122,123,124].
In this context, the knockout of the FABP3 gene or inhibition by FABP3-specific ligands ameliorates the axodendritic degeneration and neuronal loss induced by MPP+ [104] and MPTP [125,126]. Notably, the absence of FABP3 prevents the loss of mitochondrial membrane potential, abolishing the generation of stress response factors such as 4-hydroxynonenal (4-HNE) [104], which plays an essential role in the degeneration of dopaminergic neurons in Parkinson’s disease [127,128]. These observations suggest the pathogenic potential of FABP3 on the loss of mitochondrial homeostasis in Parkinson’s disease models.
Regarding other FABPs, FABP5 might also contribute to mitochondrial injury under rotenone-induced oxidative stress through its accumulation in mitochondria [129]. In addition, oligodendrocytes in Krabbe disease and the multiple sclerosis model show an abnormal localization of FABP5 in the mitochondria, leading to the formation of large pores in the mitochondrial outer membrane [130,131]. This pore formation triggers the release of inflammatory substances, leading to cell death. Furthermore, FABP3 and FABP5 are found together with 4-HNE in the mitochondria of neurons in the mice model of transient cerebral ischemia, which is also mediated by the formation of pores in the mitochondrial membrane [132,133]. These data suggest that not only FABP3 but also another subtype, FABP5, participates in mitochondrial dysfunction in the process of neuronal degeneration (Figure 3).
Glial cells are also closely involved in the pathogenesis of Parkinson’s disease [134,135]. FABP7 is involved in pathological processes in the central nervous system. Increased levels of FABP7 were observed in various parts of the hippocampus in response to cerebral ischemia in young adult monkeys [136,137]. In mouse models, FABP7 expression was up-regulated in glial fibrillary acidic protein (GFAP)-positive cells in response to spinal cord injury and autoimmune conditions [138], including experimental autoimmune encephalomyelitis [131,139]. Furthermore, FABP7 contributes to the accumulation of α-synuclein in mitochondria in oligodendrocytes to lose the membrane potential in the multiple system atrophy model mice [140,141]. These findings support the possibility of FABP7 participating in the process of neuronal degeneration.

7. Therapeutic Potential of FABP-Targeting Drugs for Parkinson’s Disease

As previously mentioned, the presence of FABPs is closely related to the maintenance of dopaminergic homeostasis and mitochondrial function in neurodegeneration. FABP3 and FABP5 are essential for the neurotoxicity expression of α-synuclein. Considering the potential for drug development targeting FABP3, we observed the absence of α-synuclein neurotoxicity in genetic knockout cells of FABP3. It was found that FABP3 can bind to α-synuclein in a 1:1 ratio, promoting α-synuclein-FABP3 aggregate formation [115]. Consequently, we attempted to create a drug that could inhibit the interaction between FABP3 and α-synuclein [126,142].
The FABP3 ligand, as a targeted drug for FABP3, inhibits α-synuclein aggregation and prevents its propagation in the brain and subsequent neuronal loss [106]. The FABP3 ligand effectively prevents the degeneration of dopaminergic neurons and restores motor dysfunction to a healthy level in a Parkinson’s disease mouse model [126]. Additionally, an α-synuclein mimicking peptide that inhibits the binding and aggregate formation of α-synuclein-FABP3 is capable of restoring memory and learning abilities in a Parkinson’s disease dementia mouse model [142]. This peptide has the potential to alleviate α-synuclein phosphorylation, which links to neurotoxicity. We are currently conducting preclinical trials of small molecules targeting FABP3 (Japan Agency for Medical Research and Development (AMED), Translational Research Program 23ym0126095h0002). Although the development of fundamental treatments for Parkinson’s disease is challenging, we aim to achieve early development and deliver it to Parkinson’s disease patients and their families.
With regard to FABP5, the specific ligand for FABP5 abolishes α-synuclein accumulation and translocation to mitochondria, which alleviates neurotoxicity in Parkinson’s disease model neuronal cells [129], as well as ameliorates oligodendrocyte injury in multiple sclerosis mouse models [131] and psychosine-induced apoptosis in Krabbe disease models [130]. Furthermore, the FABP5 ligand also has the ability to ameliorate FABP3/5-induced mitochondrial injury, including lipid peroxidation and BAX-related apoptotic signaling, indicating the potential neuroprotective treatment for ischemic stroke [132,133]. In addition, regarding FABP7, the ligand alleviates FABP7-induced α-synuclein oligomerization and aggregation, thereby rescuing glial and oligodendrocytes from cell death in multiple system atrophy model cells and mice [140,141,143]. These data suggest the promising potential of unique drug developments targeting FABP to regulate dopaminergic signaling and protect neuronal homeostasis.

8. Diagnostic Potential of FABPs as a Prodromal Biomarker for Parkinson’s Diseases

In recent years, there has been remarkable progress in targeted therapy for Parkinson’s disease, allowing for the precise alleviation of clinical symptoms [144,145]. However, challenges remain that affect the quality of life for patients, including subcutaneous device attachment, reduced medication efficacy due to disease progression, and the presence of OFF symptoms. In neurodegenerative diseases, post-onset treatment does not yield favorable prognoses, making fundamental treatment challenging. To overcome Parkinson’s disease, it is necessary to predict disease risk before onset and achieve early therapeutic interventions. Numerous excellent studies have been reported on diagnostic techniques for Parkinson’s disease using analyses of cerebrospinal fluid (CSF), serum, and plasma biomarkers targeting α-synuclein and other related proteins [146,147,148,149]. Additionally, recent studies have demonstrated that severe olfactory impairment is a useful early biomarker for Parkinson’s disease dementia [64,65,150,151,152].
Previously, certain reports have suggested methods to predict Parkinson’s disease using FABP3 as a biomarker in the cerebrospinal fluid or serum [2,32,153,154,155,156,157,158]. These investigations spotlight the pathogenic significance of the cerebrospinal fluid and serum FABP3 levels, indicating characteristic modifications in Parkinson’s disease, Parkinson’s disease dementia, and dementia with Lewy bodies. Elevated FABP3 levels are invariably observed in cerebrospinal fluid and serum specimens from patients with these disorders. Notably, higher FABP3 levels are observed in patients with dementia with Lewy bodies than in Parkinson’s disease dementia and Alzheimer’s disease, indicating the potential of FABP3 as a biomarker distinctive to dementia with Lewy bodies.
Therefore, with the aim of establishing a more accurate technique using plasma biomarkers to predict disease risk and discriminate potential disease types, we recently focused on FABP family proteins, essential for the propagation and expression of α-synuclein neurotoxicity, and developed a differential diagnostic technique for neurodegenerative disorders through a multi-marker analysis in combination with known biomarkers [31]. This method not only allows the discrimination between healthy individuals and those with Parkinson’s disease but also enables the prediction of Lewy body dementia, Alzheimer’s disease, and mild cognitive impairment. Specifically, we established a scoring technique that can differentiate Parkinson’s disease, Lewy body dementia, and Alzheimer’s disease from healthy individuals and also can distinguish between different neurodegenerative diseases (Figure 4) [31]. These multi-marker disease risk scores correlate with clinical symptoms in each disease, specifically mini-mental state examination (MMSE) scores and Hoehn–Yahr stages in Parkinson’s disease. We hope that this technology will contribute to the differentiation of clinically challenging conditions such as Parkinson’s disease dementia, Lewy body dementia, and Alzheimer’s disease in clinical settings.
In addition to ELISA analysis, metabolomics is a useful tool for measuring biomarkers [159,160]. ELISA analysis of plasma can measure the amount of a specific protein. This method is considered very accurate and reliable when the protein to be measured is known. Therefore, understanding the pathogenetic significance of the target proteins is essential. In contrast, metabolomics analysis, on the other hand, provides a comprehensive measurement of metabolites in vivo. The advantage of this method is that measurement is possible even if the metabolite to be measured is unknown. In this respect, compared to ELISA analysis, it may require more advanced techniques to account for measurement accuracy. Previous metabolomics studies successfully revealed lipid dysregulation in Parkinson’s disease [161,162]. Through utilizing the advantages of both of these approaches and establishing more accurate and highly specific biomarker predictions, it would be possible to produce a very early diagnosis of Parkinson’s disease before its onset.

9. Conclusions

In this review, we focused on FABPs and provided an overview of recent findings regarding the pathogenic impact of FABPs in the propagation and toxic expression of α-synuclein in neurodegenerative processes, as well as in the loss of mitochondrial homeostasis via VDAC-1/BAX complex formation to reduce intramitochondrial cytochrome C and membrane potential, resulting in apoptosis. FABP3 was found to be essential for the intracellular accumulation and oligomerization of α-synuclein, a pathogenic protein in Parkinson’s disease, as well as for its propagation in the brain. Additionally, FABP3 was responsible for the acceleration of TH phosphorylation and its subsequent degradation, which led to the loss of mesencephalic TH protein. FABP3 was also implicated in mitochondrial dysfunction and found to be a cause of ROS production in dopaminergic neurons.
In addition, FABP5 was also identified as essential for the toxic expression of α-synuclein through inducing apoptosis through the accumulation of α-synuclein in mitochondria and the formation of macropores on the mitochondrial membrane. Similarly, FABP7 was determined to be essential for the uptake and toxic expression of α-synuclein in glial cells, leading to apoptosis through its accumulation in mitochondria. Specific ligands for these FABPs were capable of inhibiting the formation of intracellular α-synuclein aggregates and mitochondrial injury, offering potential preventive therapeutic options for neurotoxicity.
Furthermore, the increased presence of FABP3 in the plasma of Parkinson’s disease patients indicated its potential benefit in predicting disease risk and discriminating from other α-synucleinopathies and Alzheimer’s disease, using multi-marker analyses. These findings suggest the utility of FABPs as pharmacological targets and biomarkers for Parkinson’s disease. We sincerely hope that these insights will deepen our understanding of the pathogenesis of Parkinson’s disease and contribute to the development of unique therapeutics and early diagnosis.

10. Patents

For the peptides used to prevent or treat synucleinopathy, the patent WO/2022/024693 can be referred to. Additionally, the biomarker for the diagnosis of dementia is documented in patent WO/2022/163818.

Author Contributions

Conceptualization, I.K. and K.F.; writing—original draft preparation, I.K.; writing—review and editing, I.K. and K.F.; funding acquisition, I.K. and K.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the Japan Agency for Medical Research and Development (AMED) (JP20dm0107071) to K.F. and grants from the Japan AMED (22ym0126095h0001), the Japan Society for the Promotion of Science (JSPS) KAKENHI (22K06644), and the Takeda Science Foundation to I.K.

Institutional Review Board Statement

In relation to the original data used for this review, all the animal studies were conducted in accordance with the Tohoku University institutional guidelines. Ethical approval was obtained from the Institutional Animal Care and Use Committee of the Tohoku University Environmental and Safety Committee (approval numbers 2019PhLM0-021 and 2019PhA-024). The human experiments conducted in the original study adhered to the principles outlined in the Declaration of Helsinki. The research protocols were approved by the Ethics Committees of Tohoku University (approval number PH19-5) and the National Hospital Organization Sendai Nishitaga Hospital (approval numbers 29-3 and 29-10).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to thank Emeritus Toshiharu Nagatsu of Fujita Medical University for his academic guidance on the biochemistry of tyrosine hydroxylase. We also appreciate Kazuko Hasegawa, Head of the Neuropathy Unit, National Hospital Organization Sagamihara Hospital, for the autopsy brain analysis of phosphorylated tyrosine hydroxylase. The authors would also like to thank Atsushi Takeda, Director of the National Hospital Organization Sendai Nishitaga Hospital, for supporting biomarker analysis using human plasma.

Conflicts of Interest

Author Kohji Fukunaga was employed by the company BRI Pharma Inc. The remaining authors declare that the research was conducted without any commercial relationships that could be construed as a potential conflict of interest. No commercial funding is applicable for this review article. Professor Fukunaga has established his own company. At this stage, any of the results of this academic research have not been licensed out, related to his company, nor have any profits been generated. Therefore, although his position includes BRI Pharma Inc., due to establishing a company by himself, the company itself is not involved in the academic research in this article.

References

  1. Zimmermann-Ivol, C.G.; Burkhard, P.R.; Le Floch-Rohr, J.; Allard, L.; Hochstrasser, D.F.; Sanchez, J.C. Fatty acid binding protein as a serum marker for the early diagnosis of stroke: A pilot study. Mol. Cell. Proteom. 2004, 3, 66–72. [Google Scholar] [CrossRef] [PubMed]
  2. Mollenhauer, B.; Steinacker, P.; Bahn, E.; Bibl, M.; Brechlin, P.; Schlossmacher, M.G.; Locascio, J.J.; Wiltfang, J.; Kretzschmar, H.A.; Poser, S.; et al. Serum heart-type fatty acid-binding protein and cerebrospinal fluid tau: Marker candidates for dementia with Lewy bodies. Neurodegener. Dis. 2007, 4, 366–375. [Google Scholar] [CrossRef] [PubMed]
  3. Wada-Isoe, K.; Imamura, K.; Kitamaya, M.; Kowa, H.; Nakashima, K. Serum heart-fatty acid binding protein levels in patients with Lewy body disease. J. Neurol. Sci. 2008, 266, 20–24. [Google Scholar] [CrossRef] [PubMed]
  4. Teunissen, C.E.; Veerhuis, R.; De Vente, J.; Verhey, F.R.; Vreeling, F.; van Boxtel, M.P.; Glatz, J.F.; Pelsers, M.A. Brain-specific fatty acid-binding protein is elevated in serum of patients with dementia-related diseases. Eur. J. Neurol. 2011, 18, 865–871. [Google Scholar] [CrossRef] [PubMed]
  5. Xu, H.; Diolintzi, A.; Storch, J. Fatty acid-binding proteins: Functional understanding and diagnostic implications. Curr. Opin. Clin. Nutr. Metab. Care 2019, 22, 407–412. [Google Scholar] [CrossRef] [PubMed]
  6. Ockner, R.K.; Manning, J.A.; Poppenhausen, R.B.; Ho, W.K. A binding protein for fatty acids in cytosol of intestinal mucosa, liver, myocardium, and other tissues. Science 1972, 177, 56–58. [Google Scholar] [CrossRef] [PubMed]
  7. Veerkamp, J.H. Fatty acid transport and fatty acid-binding proteins. Proc. Nutr. Soc. 1995, 54, 23–37. [Google Scholar] [CrossRef]
  8. Coe, N.R.; Bernlohr, D.A. Physiological properties and functions of intracellular fatty acid-binding proteins. Biochim. Biophys. Acta 1998, 1391, 287–306. [Google Scholar] [CrossRef]
  9. Storch, J.; Thumser, A.E. The fatty acid transport function of fatty acid-binding proteins. Biochim. Biophys. Acta 2000, 1486, 28–44. [Google Scholar] [CrossRef]
  10. Chmurzyńska, A. The multigene family of fatty acid-binding proteins (FABPs): Function, structure and polymorphism. J. Appl. Genet. 2006, 47, 39–48. [Google Scholar] [CrossRef]
  11. Smathers, R.L.; Petersen, D.R. The human fatty acid-binding protein family: Evolutionary divergences and functions. Hum. Genom. 2011, 5, 170–191. [Google Scholar] [CrossRef] [PubMed]
  12. Zimmerman, A.W.; Veerkamp, J.H. New insights into the structure and function of fatty acid-binding proteins. Cell. Mol. Life Sci. 2002, 59, 1096–1116. [Google Scholar] [CrossRef] [PubMed]
  13. Zimmerman, A.W.; van Moerkerk, H.T.; Veerkamp, J.H. Ligand specificity and conformational stability of human fatty acid-binding proteins. Int. J. Biochem. Cell Biol. 2001, 33, 865–876. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, R.Z.; Mita, R.; Beaulieu, M.; Gao, Z.; Godbout, R. Fatty acid binding proteins in brain development and disease. Int. J. Dev. Biol. 2010, 54, 1229–1239. [Google Scholar] [CrossRef] [PubMed]
  15. Xu, B.; Chen, L.; Zhan, Y.; Marquez, K.N.S.; Zhuo, L.; Qi, S.; Zhu, J.; He, Y.; Chen, X.; Zhang, H.; et al. The Biological Functions and Regulatory Mechanisms of Fatty Acid Binding Protein 5 in Various Diseases. Front. Cell Dev. Biol. 2022, 10, 857919. [Google Scholar] [CrossRef]
  16. Binas, B.; Danneberg, H.; McWhir, J.; Mullins, L.; Clark, A.J. Requirement for the heart-type fatty acid binding protein in cardiac fatty acid utilization. FASEB J. 1999, 13, 805–812. [Google Scholar] [CrossRef] [PubMed]
  17. Schaap, F.G.; Binas, B.; Danneberg, H.; van der Vusse, G.J.; Glatz, J.F. Impaired long-chain fatty acid utilization by cardiac myocytes isolated from mice lacking the heart-type fatty acid binding protein gene. Circ. Res. 1999, 85, 329–337. [Google Scholar] [CrossRef]
  18. Owada, Y.; Yoshimoto, T.; Kondo, H. Spatio-temporally differential expression of genes for three members of fatty acid binding proteins in developing and mature rat brains. J. Chem. Neuroanat. 1996, 12, 113–122. [Google Scholar] [CrossRef]
  19. Owada, Y. Fatty acid binding protein: Localization and functional significance in the brain. Tohoku J. Exp. Med. 2008, 214, 213–220. [Google Scholar] [CrossRef]
  20. Hanhoff, T.; Lucke, C.; Spener, F. Insights into binding of fatty acids by fatty acid binding proteins. Mol. Cell. Biochem. 2002, 239, 45–54. [Google Scholar] [CrossRef]
  21. Armstrong, E.H.; Goswami, D.; Griffin, P.R.; Noy, N.; Ortlund, E.A. Structural basis for ligand regulation of the fatty acid-binding protein 5, peroxisome proliferator-activated receptor beta/delta (FABP5-PPARbeta/delta) signaling pathway. J. Biol. Chem. 2014, 289, 14941–14954. [Google Scholar] [CrossRef] [PubMed]
  22. Mita, R.; Beaulieu, M.J.; Field, C.; Godbout, R. Brain fatty acid-binding protein and omega-3/omega-6 fatty acids: Mechanistic insight into malignant glioma cell migration. J. Biol. Chem. 2010, 285, 37005–37015. [Google Scholar] [CrossRef] [PubMed]
  23. Perrin, R.J.; Woods, W.S.; Clayton, D.F.; George, J.M. Exposure to long chain polyunsaturated fatty acids triggers rapid multimerization of synucleins. J. Biol. Chem. 2001, 276, 41958–41962. [Google Scholar] [CrossRef] [PubMed]
  24. Sharon, R.; Goldberg, M.S.; Bar-Josef, I.; Betensky, R.A.; Shen, J.; Selkoe, D.J. alpha-Synuclein occurs in lipid-rich high molecular weight complexes, binds fatty acids, and shows homology to the fatty acid-binding proteins. Proc. Natl. Acad. Sci. USA 2001, 98, 9110–9115. [Google Scholar] [CrossRef] [PubMed]
  25. Sharon, R.; Bar-Joseph, I.; Frosch, M.P.; Walsh, D.M.; Hamilton, J.A.; Selkoe, D.J. The formation of highly soluble oligomers of alpha-synuclein is regulated by fatty acids and enhanced in Parkinson’s disease. Neuron 2003, 37, 583–595. [Google Scholar] [CrossRef] [PubMed]
  26. Yakunin, E.; Loeb, V.; Kisos, H.; Biala, Y.; Yehuda, S.; Yaari, Y.; Selkoe, D.J.; Sharon, R. Alpha-synuclein neuropathology is controlled by nuclear hormone receptors and enhanced by docosahexaenoic acid in a mouse model for Parkinson’s disease. Brain Pathol. 2012, 22, 280–294. [Google Scholar] [CrossRef] [PubMed]
  27. Miyake, Y.; Sasaki, S.; Tanaka, K.; Fukushima, W.; Kiyohara, C.; Tsuboi, Y.; Yamada, T.; Oeda, T.; Miki, T.; Kawamura, N.; et al. Dietary fat intake and risk of Parkinson’s disease: A case-control study in Japan. J. Neurol. Sci. 2010, 288, 117–122. [Google Scholar] [CrossRef] [PubMed]
  28. Pelsers, M.M.; Hanhoff, T.; Van der Voort, D.; Arts, B.; Peters, M.; Ponds, R.; Honig, A.; Rudzinski, W.; Spener, F.; de Kruijk, J.R.; et al. Brain- and heart-type fatty acid-binding proteins in the brain: Tissue distribution and clinical utility. Clin. Chem. 2004, 50, 1568–1575. [Google Scholar] [CrossRef]
  29. Basso, M.; Giraudo, S.; Corpillo, D.; Bergamasco, B.; Lopiano, L.; Fasano, M. Proteome analysis of human substantia nigra in Parkinson’s disease. Proteomics 2004, 4, 3943–3952. [Google Scholar] [CrossRef]
  30. Oizumi, H.; Yamasaki, K.; Suzuki, H.; Hasegawa, T.; Sugimura, Y.; Baba, T.; Fukunaga, K.; Takeda, A. Fatty Acid-Binding Protein 3 Expression in the Brain and Skin in Human Synucleinopathies. Front. Aging Neurosci. 2021, 13, 648982. [Google Scholar] [CrossRef]
  31. Kawahata, I.; Sekimori, T.; Oizumi, H.; Takeda, A.; Fukunaga, K. Using Fatty Acid-Binding Proteins as Potential Biomarkers to Discriminate between Parkinson’s Disease and Dementia with Lewy Bodies: Exploration of a Novel Technique. Int. J. Mol. Sci. 2023, 24, 13267. [Google Scholar] [CrossRef] [PubMed]
  32. Backstrom, D.C.; Eriksson Domellof, M.; Linder, J.; Olsson, B.; Ohrfelt, A.; Trupp, M.; Zetterberg, H.; Blennow, K.; Forsgren, L. Cerebrospinal Fluid Patterns and the Risk of Future Dementia in Early, Incident Parkinson Disease. JAMA Neurol. 2015, 72, 1175–1182. [Google Scholar] [CrossRef] [PubMed]
  33. Ockner, R.K.; Manning, J.A.; Kane, J.P. Fatty acid binding protein. Isolation from rat liver, characterization, and immunochemical quantification. J. Biol. Chem. 1982, 257, 7872–7878. [Google Scholar] [CrossRef]
  34. Lowe, J.B.; Boguski, M.S.; Sweetser, D.A.; Elshourbagy, N.A.; Taylor, J.M.; Gordon, J.I. Human liver fatty acid binding protein. Isolation of a full length cDNA and comparative sequence analyses of orthologous and paralogous proteins. J. Biol. Chem. 1985, 260, 3413–3417. [Google Scholar] [CrossRef] [PubMed]
  35. Sweetser, D.A.; Lowe, J.B.; Gordon, J.I. The nucleotide sequence of the rat liver fatty acid-binding protein gene. Evidence that exon 1 encodes an oligopeptide domain shared by a family of proteins which bind hydrophobic ligands. J. Biol. Chem. 1986, 261, 5553–5561. [Google Scholar] [CrossRef] [PubMed]
  36. Alpers, D.H.; Strauss, A.W.; Ockner, R.K.; Bass, N.M.; Gordon, J.I. Cloning of a cDNA encoding rat intestinal fatty acid binding protein. Proc. Natl. Acad. Sci. USA 1984, 81, 313–317. [Google Scholar] [CrossRef] [PubMed]
  37. Sweetser, D.A.; Birkenmeier, E.H.; Klisak, I.J.; Zollman, S.; Sparkes, R.S.; Mohandas, T.; Lusis, A.J.; Gordon, J.I. The human and rodent intestinal fatty acid binding protein genes. A comparative analysis of their structure, expression, and linkage relationships. J. Biol. Chem. 1987, 262, 16060–16071. [Google Scholar] [CrossRef]
  38. Green, R.P.; Cohn, S.M.; Sacchettini, J.C.; Jackson, K.E.; Gordon, J.I. The mouse intestinal fatty acid binding protein gene: Nucleotide sequence, pattern of developmental and regional expression, and proposed structure of its protein product. DNA Cell Biol. 1992, 11, 31–41. [Google Scholar] [CrossRef]
  39. Sacchettini, J.C.; Said, B.; Schulz, H.; Gordon, J.I. Rat heart fatty acid-binding protein is highly homologous to the murine adipocyte 422 protein and the P2 protein of peripheral nerve myelin. J. Biol. Chem. 1986, 261, 8218–8223. [Google Scholar] [CrossRef]
  40. Heuckeroth, R.O.; Birkenmeier, E.H.; Levin, M.S.; Gordon, J.I. Analysis of the tissue-specific expression, developmental regulation, and linkage relationships of a rodent gene encoding heart fatty acid binding protein. J. Biol. Chem. 1987, 262, 9709–9717. [Google Scholar] [CrossRef]
  41. Kurtz, A.; Zimmer, A.; Schnütgen, F.; Brüning, G.; Spener, F.; Müller, T. The expression pattern of a novel gene encoding brain-fatty acid binding protein correlates with neuronal and glial cell development. Development 1994, 120, 2637–2649. [Google Scholar] [CrossRef] [PubMed]
  42. Sellner, P.A.; Chu, W.; Glatz, J.F.; Berman, N.E. Developmental role of fatty acid-binding proteins in mouse brain. Brain Res. Dev. Brain Res. 1995, 89, 33–46. [Google Scholar] [CrossRef] [PubMed]
  43. Veerkamp, J.H.; Zimmerman, A.W. Fatty acid-binding proteins of nervous tissue. J. Mol. Neurosci. 2001, 16, 133–142; discussion 151–157. [Google Scholar] [CrossRef] [PubMed]
  44. Spiegelman, B.M.; Frank, M.; Green, H. Molecular cloning of mRNA from 3T3 adipocytes. Regulation of mRNA content for glycerophosphate dehydrogenase and other differentiation-dependent proteins during adipocyte development. J. Biol. Chem. 1983, 258, 10083–10089. [Google Scholar] [CrossRef] [PubMed]
  45. Bernlohr, D.A.; Angus, C.W.; Lane, M.D.; Bolanowski, M.A.; Kelly, T.J., Jr. Expression of specific mRNAs during adipose differentiation: Identification of an mRNA encoding a homologue of myelin P2 protein. Proc. Natl. Acad. Sci. USA 1984, 81, 5468–5472. [Google Scholar] [CrossRef] [PubMed]
  46. Amri, E.Z.; Bertrand, B.; Ailhaud, G.; Grimaldi, P. Regulation of adipose cell differentiation. I. Fatty acids are inducers of the aP2 gene expression. J. Lipid Res. 1991, 32, 1449–1456. [Google Scholar] [CrossRef] [PubMed]
  47. Bernlohr, D.A.; Coe, N.R.; Simpson, M.A.; Hertzel, A.V. Regulation of gene expression in adipose cells by polyunsaturated fatty acids. Adv. Exp. Med. Biol. 1997, 422, 145–156. [Google Scholar] [CrossRef] [PubMed]
  48. Madsen, P.; Rasmussen, H.H.; Leffers, H.; Honoré, B.; Celis, J.E. Molecular cloning and expression of a novel keratinocyte protein (psoriasis-associated fatty acid-binding protein [PA-FABP]) that is highly up-regulated in psoriatic skin and that shares similarity to fatty acid-binding proteins. J. Investig. Dermatol. 1992, 99, 299–305. [Google Scholar] [CrossRef]
  49. Krieg, P.; Feil, S.; Fürstenberger, G.; Bowden, G.T. Tumor-specific overexpression of a novel keratinocyte lipid-binding protein. Identification and characterization of a cloned sequence activated during multistage carcinogenesis in mouse skin. J. Biol. Chem. 1993, 268, 17362–17369. [Google Scholar] [CrossRef]
  50. Liu, Y.; Longo, L.D.; De Leon, M. In situ and immunocytochemical localization of E-FABP mRNA and protein during neuronal migration and differentiation in the rat brain. Brain Res. 2000, 852, 16–27. [Google Scholar] [CrossRef]
  51. Gong, Y.Z.; Everett, E.T.; Schwartz, D.A.; Norris, J.S.; Wilson, F.A. Molecular cloning, tissue distribution, and expression of a 14-kDa bile acid-binding protein from rat ileal cytosol. Proc. Natl. Acad. Sci. USA 1994, 91, 4741–4745. [Google Scholar] [CrossRef]
  52. Vodenlich, A.D., Jr.; Gong, Y.Z.; Geoghegan, K.F.; Lin, M.C.; Lanzetti, A.J.; Wilson, F.A. Identification of the 14 kDa bile acid transport protein of rat ileal cytosol as gastrotropin. Biochem. Biophys. Res. Commun. 1991, 177, 1147–1154. [Google Scholar] [CrossRef]
  53. Walz, D.A.; Wider, M.D.; Snow, J.W.; Dass, C.; Desiderio, D.M. The complete amino acid sequence of porcine gastrotropin, an ileal protein which stimulates gastric acid and pepsinogen secretion. J. Biol. Chem. 1988, 263, 14189–14195. [Google Scholar] [CrossRef] [PubMed]
  54. Gantz, I.; Nothwehr, S.F.; Lucey, M.; Sacchettini, J.C.; DelValle, J.; Banaszak, L.J.; Naud, M.; Gordon, J.I.; Yamada, T. Gastrotropin: Not an enterooxyntin but a member of a family of cytoplasmic hydrophobic ligand binding proteins. J. Biol. Chem. 1989, 264, 20248–20254. [Google Scholar] [CrossRef]
  55. Feng, L.; Hatten, M.E.; Heintz, N. Brain lipid-binding protein (BLBP): A novel signaling system in the developing mammalian CNS. Neuron 1994, 12, 895–908. [Google Scholar] [CrossRef]
  56. Shimizu, F.; Watanabe, T.K.; Shinomiya, H.; Nakamura, Y.; Fujiwara, T. Isolation and expression of a cDNA for human brain fatty acid-binding protein (B-FABP). Biochim. Biophys. Acta 1997, 1354, 24–28. [Google Scholar] [CrossRef] [PubMed]
  57. Oko, R.; Morales, C.R. A novel testicular protein, with sequence similarities to a family of lipid binding proteins, is a major component of the rat sperm perinuclear theca. Dev. Biol. 1994, 166, 235–245. [Google Scholar] [CrossRef]
  58. Pouresmaeili, F.; Morales, C.R.; Oko, R. Molecular cloning and structural analysis of the gene encoding PERF 15 protein present in the perinuclear theca of the rat spermatozoa. Biol. Reprod. 1997, 57, 655–659. [Google Scholar] [CrossRef] [PubMed]
  59. de Lau, L.M.; Breteler, M.M. Epidemiology of Parkinson’s disease. Lancet Neurol. 2006, 5, 525–535. [Google Scholar] [CrossRef]
  60. Hirsch, L.; Jette, N.; Frolkis, A.; Steeves, T.; Pringsheim, T. The Incidence of Parkinson’s Disease: A Systematic Review and Meta-Analysis. Neuroepidemiology 2016, 46, 292–300. [Google Scholar] [CrossRef]
  61. Ruan, X.; Lin, F.; Wu, D.; Chen, L.; Weng, H.; Yu, J.; Wang, Y.; Chen, Y.; Chen, X.; Ye, Q.; et al. Comparative Efficacy and Safety of Dopamine Agonists in Advanced Parkinson’s Disease With Motor Fluctuations: A Systematic Review and Network Meta-Analysis of Double-Blind Randomized Controlled Trials. Front. Neurosci. 2021, 15, 728083. [Google Scholar] [CrossRef]
  62. Parkinson, J. An essay on the shaking palsy. 1817. J. Neuropsychiatry Clin. Neurosci. 2002, 14, 223–236; discussion 222. [Google Scholar] [CrossRef] [PubMed]
  63. Phillips, O.; Ghosh, D.; Fernandez, H.H. Parkinson Disease Dementia Management: An Update of Current Evidence and Future Directions. Curr. Treat. Options Neurol. 2023, 25, 93–119. [Google Scholar] [CrossRef]
  64. Takeda, A.; Baba, T.; Kikuchi, A.; Hasegawa, T.; Sugeno, N.; Konno, M.; Miura, E.; Mori, E. Olfactory dysfunction and dementia in Parkinson’s disease. J. Park. Dis. 2014, 4, 181–187. [Google Scholar] [CrossRef] [PubMed]
  65. Takeda, A. How Useful is an Olfactory Test for Diagnosing Alzheimer’s Syndrome? Brain Nerve 2023, 75, 943–948. [Google Scholar] [CrossRef] [PubMed]
  66. Rey, N.L.; Wesson, D.W.; Brundin, P. The olfactory bulb as the entry site for prion-like propagation in neurodegenerative diseases. Neurobiol. Dis. 2018, 109, 226–248. [Google Scholar] [CrossRef] [PubMed]
  67. Spillantini, M.G.; Schmidt, M.L.; Lee, V.M.; Trojanowski, J.Q.; Jakes, R.; Goedert, M. Alpha-synuclein in Lewy bodies. Nature 1997, 388, 839–840. [Google Scholar] [CrossRef]
  68. Spillantini, M.G.; Crowther, R.A.; Jakes, R.; Hasegawa, M.; Goedert, M. alpha-Synuclein in filamentous inclusions of Lewy bodies from Parkinson’s disease and dementia with lewy bodies. Proc. Natl. Acad. Sci. USA 1998, 95, 6469–6473. [Google Scholar] [CrossRef] [PubMed]
  69. Spillantini, M.G.; Crowther, R.A.; Jakes, R.; Cairns, N.J.; Lantos, P.L.; Goedert, M. Filamentous alpha-synuclein inclusions link multiple system atrophy with Parkinson’s disease and dementia with Lewy bodies. Neurosci. Lett. 1998, 251, 205–208. [Google Scholar] [CrossRef]
  70. Shetty, A.S.; Bhatia, K.P.; Lang, A.E. Dystonia and Parkinson’s disease: What is the relationship? Neurobiol. Dis. 2019, 132, 104462. [Google Scholar] [CrossRef]
  71. Montague, P.R.; Hyman, S.E.; Cohen, J.D. Computational roles for dopamine in behavioural control. Nature 2004, 431, 760–767. [Google Scholar] [CrossRef] [PubMed]
  72. Schultz, W. Behavioral theories and the neurophysiology of reward. Annu. Rev. Psychol. 2006, 57, 87–115. [Google Scholar] [CrossRef] [PubMed]
  73. Berridge, K.C. From prediction error to incentive salience: Mesolimbic computation of reward motivation. Eur. J. Neurosci. 2012, 35, 1124–1143. [Google Scholar] [CrossRef]
  74. Stern, G. The effects of lesions in the substantia nigra. Brain 1966, 89, 449–478. [Google Scholar] [CrossRef] [PubMed]
  75. Hoefer, P.F.A. Action Potentials of Muscles in Rigidity and Tremor. Arch. Neurol. Psychiatry 1940, 43, 704–725. [Google Scholar] [CrossRef]
  76. Nagatsu, T. The catecholamine system in health and disease -Relation to tyrosine 3-monooxygenase and other catecholamine-synthesizing enzymes. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2007, 82, 388–415. [Google Scholar] [CrossRef] [PubMed]
  77. Nagatsu, T.; Nagatsu, I. Tyrosine hydroxylase (TH), its cofactor tetrahydrobiopterin (BH4), other catecholamine-related enzymes, and their human genes in relation to the drug and gene therapies of Parkinson’s disease (PD): Historical overview and future prospects. J. Neural Transm. 2016, 123, 1255–1278. [Google Scholar] [CrossRef] [PubMed]
  78. Nagatsu, T. Catecholamines and Parkinson’s disease: Tyrosine hydroxylase (TH) over tetrahydrobiopterin (BH4) and GTP cyclohydrolase I (GCH1) to cytokines, neuromelanin, and gene therapy: A historical overview. J. Neural Transm. 2023. ahead of print. [Google Scholar] [CrossRef]
  79. Nagatsu, T.; Levitt, M.; Udenfriend, S. Tyrosine Hydroxylase. The initial step in norepinephrine biosynthesis. J. Biol. Chem. 1964, 239, 2910–2917. [Google Scholar] [CrossRef]
  80. Klein, M.O.; Battagello, D.S.; Cardoso, A.R.; Hauser, D.N.; Bittencourt, J.C.; Correa, R.G. Dopamine: Functions, Signaling, and Association with Neurological Diseases. Cell. Mol. Neurobiol. 2019, 39, 31–59. [Google Scholar] [CrossRef]
  81. Helman, G.; Pappa, M.B.; Pearl, P.L. Widening Phenotypic Spectrum of AADC Deficiency, a Disorder of Dopamine and Serotonin Synthesis. JIMD Rep. 2014, 17, 23–27. [Google Scholar] [CrossRef]
  82. Atwal, P.S.; Donti, T.R.; Cardon, A.L.; Bacino, C.A.; Sun, Q.; Emrick, L.; Reid Sutton, V.; Elsea, S.H. Aromatic L-amino acid decarboxylase deficiency diagnosed by clinical metabolomic profiling of plasma. Mol. Genet. Metab. 2015, 115, 91–94. [Google Scholar] [CrossRef] [PubMed]
  83. Ichinose, H.; Ohye, T.; Takahashi, E.; Seki, N.; Hori, T.; Segawa, M.; Nomura, Y.; Endo, K.; Tanaka, H.; Tsuji, S.; et al. Hereditary progressive dystonia with marked diurnal fluctuation caused by mutations in the GTP cyclohydrolase I gene. Nat. Genet. 1994, 8, 236–242. [Google Scholar] [CrossRef] [PubMed]
  84. Segawa, M. Dopa-responsive dystonia. Handb. Clin. Neurol. 2011, 100, 539–557. [Google Scholar] [CrossRef] [PubMed]
  85. Mitchell, J.P.; Hardie, D.G.; Vulliet, P.R. Site-specific phosphorylation of tyrosine hydroxylase after KCl depolarization and nerve growth factor treatment of PC12 cells. J. Biol. Chem. 1990, 265, 22358–22364. [Google Scholar] [CrossRef]
  86. Campbell, D.G.; Hardie, D.G.; Vulliet, P.R. Identification of four phosphorylation sites in the N-terminal region of tyrosine hydroxylase. J. Biol. Chem. 1986, 261, 10489–10492. [Google Scholar] [CrossRef]
  87. Dunkley, P.R.; Bobrovskaya, L.; Graham, M.E.; von Nagy-Felsobuki, E.I.; Dickson, P.W. Tyrosine hydroxylase phosphorylation: Regulation and consequences. J. Neurochem. 2004, 91, 1025–1043. [Google Scholar] [CrossRef] [PubMed]
  88. Daubner, S.C.; Le, T.; Wang, S. Tyrosine hydroxylase and regulation of dopamine synthesis. Arch. Biochem. Biophys. 2011, 508, 1–12. [Google Scholar] [CrossRef]
  89. Leal, R.B.; Sim, A.T.; Goncalves, C.A.; Dunkley, P.R. Tyrosine hydroxylase dephosphorylation by protein phosphatase 2A in bovine adrenal chromaffin cells. Neurochem. Res. 2002, 27, 207–213. [Google Scholar] [CrossRef]
  90. Bezem, M.T.; Baumann, A.; Skjærven, L.; Meyer, R.; Kursula, P.; Martinez, A.; Flydal, M.I. Stable preparations of tyrosine hydroxylase provide the solution structure of the full-length enzyme. Sci. Rep. 2016, 6, 30390. [Google Scholar] [CrossRef]
  91. Bueno-Carrasco, M.T.; Cuéllar, J.; Flydal, M.I.; Santiago, C.; Kråkenes, T.A.; Kleppe, R.; López-Blanco, J.R.; Marcilla, M.; Teigen, K.; Alvira, S.; et al. Structural mechanism for tyrosine hydroxylase inhibition by dopamine and reactivation by Ser40 phosphorylation. Nat. Commun. 2022, 13, 74. [Google Scholar] [CrossRef]
  92. Urano, F.; Hayashi, N.; Arisaka, F.; Kurita, H.; Murata, S.; Ichinose, H. Molecular mechanism for pterin-mediated inactivation of tyrosine hydroxylase: Formation of insoluble aggregates of tyrosine hydroxylase. J. Biochem. 2006, 139, 625–635. [Google Scholar] [CrossRef] [PubMed]
  93. Baumann, A.; Jorge-Finnigan, A.; Jung-Kc, K.; Sauter, A.; Horvath, I.; Morozova-Roche, L.A.; Martinez, A. Tyrosine Hydroxylase Binding to Phospholipid Membranes Prompts Its Amyloid Aggregation and Compromises Bilayer Integrity. Sci. Rep. 2016, 6, 39488. [Google Scholar] [CrossRef] [PubMed]
  94. McNaught, K.S.; Belizaire, R.; Jenner, P.; Olanow, C.W.; Isacson, O. Selective loss of 20S proteasome alpha-subunits in the substantia nigra pars compacta in Parkinson’s disease. Neurosci. Lett. 2002, 326, 155–158. [Google Scholar] [CrossRef]
  95. McNaught, K.S.; Jenner, P. Proteasomal function is impaired in substantia nigra in Parkinson’s disease. Neurosci. Lett. 2001, 297, 191–194. [Google Scholar] [CrossRef] [PubMed]
  96. Zhang, N.Y.; Tang, Z.; Liu, C.W. alpha-Synuclein protofibrils inhibit 26 S proteasome-mediated protein degradation: Understanding the cytotoxicity of protein protofibrils in neurodegenerative disease pathogenesis. J. Biol. Chem. 2008, 283, 20288–20298. [Google Scholar] [CrossRef] [PubMed]
  97. Kawahata, I.; Tokuoka, H.; Parvez, H.; Ichinose, H. Accumulation of phosphorylated tyrosine hydroxylase into insoluble protein aggregates by inhibition of an ubiquitin-proteasome system in PC12D cells. J. Neural Transm. 2009, 116, 1571–1578. [Google Scholar] [CrossRef] [PubMed]
  98. Kawahata, I.; Yagishita, S.; Hasegawa, K.; Nagatsu, I.; Nagatsu, T.; Ichinose, H. Immunohistochemical analyses of the postmortem human brains from patients with Parkinson’s disease with anti-tyrosine hydroxylase antibodies. Biog. Amines 2009, 23, 1–7. [Google Scholar]
  99. Kawahata, I.; Ohtaku, S.; Tomioka, Y.; Ichinose, H.; Yamakuni, T. Dopamine or biopterin deficiency potentiates phosphorylation at (40)Ser and ubiquitination of tyrosine hydroxylase to be degraded by the ubiquitin proteasome system. Biochem. Biophys. Res. Commun. 2015, 465, 53–58. [Google Scholar] [CrossRef]
  100. Døskeland, A.P.; Flatmark, T. Ubiquitination of soluble and membrane-bound tyrosine hydroxylase and degradation of the soluble form. Eur. J. Biochem. 2002, 269, 1561–1569. [Google Scholar] [CrossRef]
  101. Kawahata, I.; Fukunaga, K. Degradation of Tyrosine Hydroxylase by the Ubiquitin-Proteasome System in the Pathogenesis of Parkinson’s Disease and Dopa-Responsive Dystonia. Int. J. Mol. Sci. 2020, 21, 3779. [Google Scholar] [CrossRef]
  102. Wu, B.; Liu, Q.; Duan, C.; Li, Y.; Yu, S.; Chan, P.; Ueda, K.; Yang, H. Phosphorylation of alpha-synuclein upregulates tyrosine hydroxylase activity in MN9D cells. Acta Histochem. 2011, 113, 32–35. [Google Scholar] [CrossRef]
  103. Canerina-Amaro, A.; Pereda, D.; Diaz, M.; Rodriguez-Barreto, D.; Casañas-Sánchez, V.; Heffer, M.; Garcia-Esparcia, P.; Ferrer, I.; Puertas-Avendaño, R.; Marin, R. Differential Aggregation and Phosphorylation of Alpha Synuclein in Membrane Compartments Associated With Parkinson Disease. Front. Neurosci. 2019, 13, 382. [Google Scholar] [CrossRef] [PubMed]
  104. Kawahata, I.; Bousset, L.; Melki, R.; Fukunaga, K. Fatty Acid-Binding Protein 3 is Critical for alpha-Synuclein Uptake and MPP(+)-Induced Mitochondrial Dysfunction in Cultured Dopaminergic Neurons. Int. J. Mol. Sci. 2019, 20, 5358. [Google Scholar] [CrossRef] [PubMed]
  105. Kawahata, I.; Sekimori, T.; Wang, H.; Wang, Y.; Sasaoka, T.; Bousset, L.; Melki, R.; Mizobata, T.; Kawata, Y.; Fukunaga, K. Dopamine D2 Long Receptors Are Critical for Caveolae-Mediated alpha-Synuclein Uptake in Cultured Dopaminergic Neurons. Biomedicines 2021, 9, 49. [Google Scholar] [CrossRef] [PubMed]
  106. Yabuki, Y.; Matsuo, K.; Kawahata, I.; Fukui, N.; Mizobata, T.; Kawata, Y.; Owada, Y.; Shioda, N.; Fukunaga, K. Fatty Acid Binding Protein 3 Enhances the Spreading and Toxicity of alpha-Synuclein in Mouse Brain. Int. J. Mol. Sci. 2020, 21, 2230. [Google Scholar] [CrossRef] [PubMed]
  107. Matsuo, K.; Kawahata, I.; Melki, R.; Bousset, L.; Owada, Y.; Fukunaga, K. Suppression of alpha-synuclein propagation after intrastriatal injection in FABP3 null mice. Brain Res. 2021, 1760, 147383. [Google Scholar] [CrossRef] [PubMed]
  108. Zhao, F.; Cheng, Z.; Piao, J.; Cui, R.; Li, B. Dopamine Receptors: Is It Possible to Become a Therapeutic Target for Depression? Front. Pharmacol. 2022, 13, 947785. [Google Scholar] [CrossRef] [PubMed]
  109. Takeuchi, Y.; Fukunaga, K. Differential subcellular localization of two dopamine D2 receptor isoforms in transfected NG108-15 cells. J. Neurochem. 2003, 85, 1064–1074. [Google Scholar] [CrossRef]
  110. Sharma, M.; Celver, J.; Octeau, J.C.; Kovoor, A. Plasma membrane compartmentalization of D2 dopamine receptors. J. Biol. Chem. 2013, 288, 12554–12568. [Google Scholar] [CrossRef]
  111. Cho, D.I.; Min, C.; Jung, K.S.; Cheong, S.Y.; Zheng, M.; Cheong, S.J.; Oak, M.H.; Cheong, J.H.; Lee, B.K.; Kim, K.M. The N-terminal region of the dopamine D2 receptor, a rhodopsin-like GPCR, regulates correct integration into the plasma membrane and endocytic routes. Br. J. Pharmacol. 2012, 166, 659–675. [Google Scholar] [CrossRef]
  112. Farde, L.; Hall, H.; Ehrin, E.; Sedvall, G. Quantitative analysis of D2 dopamine receptor binding in the living human brain by PET. Science 1986, 231, 258–261. [Google Scholar] [CrossRef] [PubMed]
  113. Martel, J.C.; Gatti McArthur, S. Dopamine Receptor Subtypes, Physiology and Pharmacology: New Ligands and Concepts in Schizophrenia. Front. Pharmacol. 2020, 11, 1003. [Google Scholar] [CrossRef] [PubMed]
  114. Uhlen, M.; Fagerberg, L.; Hallstrom, B.M.; Lindskog, C.; Oksvold, P.; Mardinoglu, A.; Sivertsson, A.; Kampf, C.; Sjostedt, E.; Asplund, A.; et al. Proteomics. Tissue-based map of the human proteome. Science 2015, 347, 1260419. [Google Scholar] [CrossRef] [PubMed]
  115. Fukui, N.; Yamamoto, H.; Miyabe, M.; Aoyama, Y.; Hongo, K.; Mizobata, T.; Kawahata, I.; Yabuki, Y.; Shinoda, Y.; Fukunaga, K.; et al. An alpha-synuclein decoy peptide prevents cytotoxic alpha-synuclein aggregation caused by fatty acid binding protein 3. J. Biol. Chem. 2021, 296, 100663. [Google Scholar] [CrossRef] [PubMed]
  116. Borsche, M.; Pereira, S.L.; Klein, C.; Grünewald, A. Mitochondria and Parkinson’s Disease: Clinical, Molecular, and Translational Aspects. J. Park. Dis. 2021, 11, 45–60. [Google Scholar] [CrossRef] [PubMed]
  117. Przedborski, S.; Jackson-Lewis, V.; Naini, A.B.; Jakowec, M.; Petzinger, G.; Miller, R.; Akram, M. The parkinsonian toxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP): A technical review of its utility and safety. J. Neurochem. 2001, 76, 1265–1274. [Google Scholar] [CrossRef] [PubMed]
  118. Davis, G.C.; Williams, A.C.; Markey, S.P.; Ebert, M.H.; Caine, E.D.; Reichert, C.M.; Kopin, I.J. Chronic Parkinsonism secondary to intravenous injection of meperidine analogues. Psychiatry Res. 1979, 1, 249–254. [Google Scholar] [CrossRef]
  119. Langston, J.W.; Ballard, P.; Tetrud, J.W.; Irwin, I. Chronic Parkinsonism in humans due to a product of meperidine-analog synthesis. Science 1983, 219, 979–980. [Google Scholar] [CrossRef]
  120. Jenner, P.; Marsden, C.D. The actions of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine in animals as a model of Parkinson’s disease. J. Neural Transm. Suppl. 1986, 20, 11–39. [Google Scholar]
  121. Heikkila, R.E.; Nicklas, W.J.; Vyas, I.; Duvoisin, R.C. Dopaminergic toxicity of rotenone and the 1-methyl-4-phenylpyridinium ion after their stereotaxic administration to rats: Implication for the mechanism of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine toxicity. Neurosci. Lett. 1985, 62, 389–394. [Google Scholar] [CrossRef]
  122. Johnson, M.E.; Bobrovskaya, L. An update on the rotenone models of Parkinson’s disease: Their ability to reproduce the features of clinical disease and model gene-environment interactions. Neurotoxicology 2015, 46, 101–116. [Google Scholar] [CrossRef] [PubMed]
  123. Innos, J.; Hickey, M.A. Using Rotenone to Model Parkinson’s Disease in Mice: A Review of the Role of Pharmacokinetics. Chem. Res. Toxicol. 2021, 34, 1223–1239. [Google Scholar] [CrossRef] [PubMed]
  124. Landau, R.; Halperin, R.; Sullivan, P.; Zibly, Z.; Leibowitz, A.; Goldstein, D.S.; Sharabi, Y. The rat rotenone model reproduces the abnormal pattern of central catecholamine metabolism found in Parkinson’s disease. Dis. Model. Mech. 2022, 15, dmm049082. [Google Scholar] [CrossRef] [PubMed]
  125. Haga, H.; Yamada, R.; Izumi, H.; Shinoda, Y.; Kawahata, I.; Miyachi, H.; Fukunaga, K. Novel fatty acid-binding protein 3 ligand inhibits dopaminergic neuronal death and improves motor and cognitive impairments in Parkinson’s disease model mice. Pharmacol. Biochem. Behav. 2020, 191, 172891. [Google Scholar] [CrossRef] [PubMed]
  126. Wang, H.; Fukunaga, K.; Cheng, A.; Wang, Y.; Arimura, N.; Yoshino, H.; Sasaki, T.; Kawahata, I. Novel FABP3 ligand, HY-11-9, ameliorates neuropathological deficits in MPTP-induced Parkinsonism in mice. J. Pharmacol. Sci. 2023, 152, 30–38. [Google Scholar] [CrossRef]
  127. Puspita, L.; Chung, S.Y.; Shim, J.W. Oxidative stress and cellular pathologies in Parkinson’s disease. Mol. Brain 2017, 10, 53. [Google Scholar] [CrossRef] [PubMed]
  128. Dias, V.; Junn, E.; Mouradian, M.M. The role of oxidative stress in Parkinson’s disease. J. Park. Dis. 2013, 3, 461–491. [Google Scholar] [CrossRef]
  129. Wang, Y.; Shinoda, Y.; Cheng, A.; Kawahata, I.; Fukunaga, K. Epidermal Fatty Acid-Binding Protein 5 (FABP5) Involvement in Alpha-Synuclein-Induced Mitochondrial Injury under Oxidative Stress. Biomedicines 2021, 9, 110. [Google Scholar] [CrossRef]
  130. Cheng, A.; Kawahata, I.; Fukunaga, K. Fatty Acid Binding Protein 5 Mediates Cell Death by Psychosine Exposure through Mitochondrial Macropores Formation in Oligodendrocytes. Biomedicines 2020, 8, 635. [Google Scholar] [CrossRef]
  131. Cheng, A.; Jia, W.; Kawahata, I.; Fukunaga, K. A novel fatty acid-binding protein 5 and 7 inhibitor ameliorates oligodendrocyte injury in multiple sclerosis mouse models. EBioMedicine 2021, 72, 103582. [Google Scholar] [CrossRef]
  132. Guo, Q.; Kawahata, I.; Degawa, T.; Ikeda-Matsuo, Y.; Sun, M.; Han, F.; Fukunaga, K. Fatty Acid-Binding Proteins Aggravate Cerebral Ischemia-Reperfusion Injury in Mice. Biomedicines 2021, 9, 529. [Google Scholar] [CrossRef] [PubMed]
  133. Guo, Q.; Kawahata, I.; Cheng, A.; Wang, H.; Jia, W.; Yoshino, H.; Fukunaga, K. Fatty acid-binding proteins 3 and 5 are involved in the initiation of mitochondrial damage in ischemic neurons. Redox Biol. 2023, 59, 102547. [Google Scholar] [CrossRef] [PubMed]
  134. de Araújo, F.M.; Cuenca-Bermejo, L.; Fernández-Villalba, E.; Costa, S.L.; Silva, V.D.A.; Herrero, M.T. Role of Microgliosis and NLRP3 Inflammasome in Parkinson’s Disease Pathogenesis and Therapy. Cell. Mol. Neurobiol. 2022, 42, 1283–1300. [Google Scholar] [CrossRef] [PubMed]
  135. MacMahon Copas, A.N.; McComish, S.F.; Fletcher, J.M.; Caldwell, M.A. The Pathogenesis of Parkinson’s Disease: A Complex Interplay Between Astrocytes, Microglia, and T Lymphocytes? Front. Neurol. 2021, 12, 666737. [Google Scholar] [CrossRef]
  136. Ma, D.; Zhang, M.; Mori, Y.; Yao, C.; Larsen, C.P.; Yamashima, T.; Zhou, L. Cellular localization of epidermal-type and brain-type fatty acid-binding proteins in adult hippocampus and their response to cerebral ischemia. Hippocampus 2010, 20, 811–819. [Google Scholar] [CrossRef] [PubMed]
  137. Boneva, N.B.; Kaplamadzhiev, D.B.; Sahara, S.; Kikuchi, H.; Pyko, I.V.; Kikuchi, M.; Tonchev, A.B.; Yamashima, T. Expression of fatty acid-binding proteins in adult hippocampal neurogenic niche of postischemic monkeys. Hippocampus 2011, 21, 162–171. [Google Scholar] [CrossRef] [PubMed]
  138. White, R.E.; McTigue, D.M.; Jakeman, L.B. Regional heterogeneity in astrocyte responses following contusive spinal cord injury in mice. J. Comp. Neurol. 2010, 518, 1370–1390. [Google Scholar] [CrossRef]
  139. Bannerman, P.; Hahn, A.; Soulika, A.; Gallo, V.; Pleasure, D. Astrogliosis in EAE spinal cord: Derivation from radial glia, and relationships to oligodendroglia. Glia 2007, 55, 57–64. [Google Scholar] [CrossRef]
  140. Cheng, A.; Wang, Y.F.; Shinoda, Y.; Kawahata, I.; Yamamoto, T.; Jia, W.B.; Yamamoto, H.; Mizobata, T.; Kawata, Y.; Fukunaga, K. Fatty acid-binding protein 7 triggers alpha-synuclein oligomerization in glial cells and oligodendrocytes associated with oxidative stress. Acta Pharmacol. Sin. 2022, 43, 552–562. [Google Scholar] [CrossRef]
  141. Cheng, A.; Kawahata, I.; Wang, Y.; Jia, W.; Wang, H.; Sekimori, T.; Chen, Y.; Suzuki, H.; Takeda, A.; Stefanova, N.; et al. Epsin2, a novel target for multiple system atrophy therapy via alpha-synuclein/FABP7 propagation. Brain 2023, 146, 3172–3180. [Google Scholar] [CrossRef]
  142. Guo, Q.; Kawahata, I.; Jia, W.; Wang, H.; Cheng, A.; Yabuki, Y.; Shioda, N.; Fukunaga, K. alpha-Synuclein decoy peptide protects mice against alpha-synuclein-induced memory loss. CNS Neurosci. Ther. 2023, 29, 1547–1560. [Google Scholar] [CrossRef]
  143. Cheng, A.; Jia, W.; Finkelstein, D.I.; Stefanova, N.; Wang, H.; Sasaki, T.; Kawahata, I.; Fukunaga, K. Pharmacological inhibition of FABP7 by MF 6 counteracts cerebellum dysfunction in an experimental multiple system atrophy mouse model. Acta Pharmacol. Sin. 2023. ahead of print. [Google Scholar] [CrossRef]
  144. Ntetsika, T.; Papathoma, P.E.; Markaki, I. Novel targeted therapies for Parkinson’s disease. Mol. Med. 2021, 27, 17. [Google Scholar] [CrossRef] [PubMed]
  145. Prasad, E.M.; Hung, S.Y. Current Therapies in Clinical Trials of Parkinson’s Disease: A 2021 Update. Pharmaceuticals 2021, 14, 717. [Google Scholar] [CrossRef] [PubMed]
  146. Mollenhauer, B.; Cullen, V.; Kahn, I.; Krastins, B.; Outeiro, T.F.; Pepivani, I.; Ng, J.; Schulz-Schaeffer, W.; Kretzschmar, H.A.; McLean, P.J.; et al. Direct quantification of CSF alpha-synuclein by ELISA and first cross-sectional study in patients with neurodegeneration. Exp. Neurol. 2008, 213, 315–325. [Google Scholar] [CrossRef] [PubMed]
  147. Li, X.; Fan, X.; Yang, H.; Liu, Y. Review of Metabolomics-Based Biomarker Research for Parkinson’s Disease. Mol. Neurobiol. 2022, 59, 1041–1057. [Google Scholar] [CrossRef] [PubMed]
  148. Ma, Z.L.; Wang, Z.L.; Zhang, F.Y.; Liu, H.X.; Mao, L.H.; Yuan, L. Biomarkers of Parkinson’s Disease: From Basic Research to Clinical Practice. Aging Dis. 2023. ahead of print. [Google Scholar] [CrossRef]
  149. Tsamourgelis, A.; Swann, P.; Chouliaras, L.; O’Brien, J.T. From protein biomarkers to proteomics in dementia with Lewy Bodies. Ageing Res. Rev. 2023, 83, 101771. [Google Scholar] [CrossRef]
  150. Baba, T.; Kikuchi, A.; Hirayama, K.; Nishio, Y.; Hosokai, Y.; Kanno, S.; Hasegawa, T.; Sugeno, N.; Konno, M.; Suzuki, K.; et al. Severe olfactory dysfunction is a prodromal symptom of dementia associated with Parkinson’s disease: A 3 year longitudinal study. Brain 2012, 135, 161–169. [Google Scholar] [CrossRef]
  151. Domellof, M.E.; Lundin, K.F.; Edstrom, M.; Forsgren, L. Olfactory dysfunction and dementia in newly diagnosed patients with Parkinson’s disease. Park. Relat. Disord. 2017, 38, 41–47. [Google Scholar] [CrossRef]
  152. Fullard, M.E.; Morley, J.F.; Duda, J.E. Olfactory Dysfunction as an Early Biomarker in Parkinson’s Disease. Neurosci. Bull. 2017, 33, 515–525. [Google Scholar] [CrossRef]
  153. Chiasserini, D.; Biscetti, L.; Eusebi, P.; Salvadori, N.; Frattini, G.; Simoni, S.; De Roeck, N.; Tambasco, N.; Stoops, E.; Vanderstichele, H.; et al. Differential role of CSF fatty acid binding protein 3, alpha-synuclein, and Alzheimer’s disease core biomarkers in Lewy body disorders and Alzheimer’s dementia. Alzheimers Res. Ther. 2017, 9, 52. [Google Scholar] [CrossRef]
  154. Jellinger, K.A.; Korczyn, A.D. Are dementia with Lewy bodies and Parkinson’s disease dementia the same disease? BMC Med. 2018, 16, 34. [Google Scholar] [CrossRef] [PubMed]
  155. Sepe, F.N.; Chiasserini, D.; Parnetti, L. Role of FABP3 as biomarker in Alzheimer’s disease and synucleinopathies. Future Neurol. 2018, 13, 199–207. [Google Scholar] [CrossRef]
  156. Sezgin, M.; Bilgic, B.; Tinaz, S.; Emre, M. Parkinson’s Disease Dementia and Lewy Body Disease. Semin. Neurol. 2019, 39, 274–282. [Google Scholar] [CrossRef] [PubMed]
  157. Walker, L.; Stefanis, L.; Attems, J. Clinical and neuropathological differences between Parkinson’s disease, Parkinson’s disease dementia and dementia with Lewy bodies—Current issues and future directions. J. Neurochem. 2019, 150, 467–474. [Google Scholar] [CrossRef] [PubMed]
  158. Parnetti, L.; Gaetani, L.; Eusebi, P.; Paciotti, S.; Hansson, O.; El-Agnaf, O.; Mollenhauer, B.; Blennow, K.; Calabresi, P. CSF and blood biomarkers for Parkinson’s disease. Lancet Neurol. 2019, 18, 573–586. [Google Scholar] [CrossRef] [PubMed]
  159. Triozzi, P.L.; Stirling, E.R.; Song, Q.; Westwood, B.; Kooshki, M.; Forbes, M.E.; Holbrook, B.C.; Cook, K.L.; Alexander-Miller, M.A.; Miller, L.D.; et al. Circulating Immune Bioenergetic, Metabolic, and Genetic Signatures Predict Melanoma Patients’ Response to Anti-PD-1 Immune Checkpoint Blockade. Clin. Cancer Res. 2022, 28, 1192–1202. [Google Scholar] [CrossRef] [PubMed]
  160. Lu, Z.; Priya Rajan, S.A.; Song, Q.; Zhao, Y.; Wan, M.; Aleman, J.; Skardal, A.; Bishop, C.; Atala, A.; Lu, B. 3D scaffold-free microlivers with drug metabolic function generated by lineage-reprogrammed hepatocytes from human fibroblasts. Biomaterials 2021, 269, 120668. [Google Scholar] [CrossRef] [PubMed]
  161. Sinclair, E.; Trivedi, D.K.; Sarkar, D.; Walton-Doyle, C.; Milne, J.; Kunath, T.; Rijs, A.M.; de Bie, R.M.A.; Goodacre, R.; Silverdale, M.; et al. Metabolomics of sebum reveals lipid dysregulation in Parkinson’s disease. Nat. Commun. 2021, 12, 1592. [Google Scholar] [CrossRef]
  162. Oizumi, H.; Sugimura, Y.; Totsune, T.; Kawasaki, I.; Ohshiro, S.; Baba, T.; Kimpara, T.; Sakuma, H.; Hasegawa, T.; Kawahata, I.; et al. Plasma sphingolipid abnormalities in neurodegenerative diseases. PLoS ONE 2022, 17, e0279315. [Google Scholar] [CrossRef]
Figure 1. Degradation or aggregation of phosphorylated TH at serine 40 residue in the Parkinson’s disease model cells. (A) The dopamine-deficient state facilitates the phosphorylation of TH, which results in the degradation or aggregation of the protein [97,99]. (B) The schematic diagram of the TH degradation pathway in the Parkinson’s disease model cells [99,100,101]. pTH: TH phosphorylated at the serine 40; PKA: cAMP-dependent protein kinase; MSK1: mitogen- and stress-activated protein kinase 1; MAPKAPK: mitogen-activated protein kinase-activated protein kinase.
Figure 1. Degradation or aggregation of phosphorylated TH at serine 40 residue in the Parkinson’s disease model cells. (A) The dopamine-deficient state facilitates the phosphorylation of TH, which results in the degradation or aggregation of the protein [97,99]. (B) The schematic diagram of the TH degradation pathway in the Parkinson’s disease model cells [99,100,101]. pTH: TH phosphorylated at the serine 40; PKA: cAMP-dependent protein kinase; MSK1: mitogen- and stress-activated protein kinase 1; MAPKAPK: mitogen-activated protein kinase-activated protein kinase.
Ijms 24 17037 g001
Figure 2. Schematic diagram of the involvement of FABP3 in the uptake and oligomerization of α-synuclein in neuronal cells. (A) FABP3 is scaffolded by dopamine D2 long type (D2L) receptors and is distributed near the plasma membrane, preferentially in the membrane rafts. FABP3 is critical for the α-synuclein accumulation and oligomerization to form aggregates [104,105,106,115]. (B) The absence of FABP3 abolishes α-synuclein accumulation and aggregation [104,105,106,107]. α-Synuclein that enters via the normal endocytic pathway is either degraded by endosomes or rapidly re-released via exocytosis. Note that there are various other mechanisms of α-synuclein uptake and possibilities for α-synuclein to enter the neuronal cells.
Figure 2. Schematic diagram of the involvement of FABP3 in the uptake and oligomerization of α-synuclein in neuronal cells. (A) FABP3 is scaffolded by dopamine D2 long type (D2L) receptors and is distributed near the plasma membrane, preferentially in the membrane rafts. FABP3 is critical for the α-synuclein accumulation and oligomerization to form aggregates [104,105,106,115]. (B) The absence of FABP3 abolishes α-synuclein accumulation and aggregation [104,105,106,107]. α-Synuclein that enters via the normal endocytic pathway is either degraded by endosomes or rapidly re-released via exocytosis. Note that there are various other mechanisms of α-synuclein uptake and possibilities for α-synuclein to enter the neuronal cells.
Ijms 24 17037 g002
Figure 3. Schematic diagram of the involvement of FABP3 and FABP5 in mitochondrial dysfunction in neuronal cells of Lewy body disease model. Neurons express FABP3 and FABP5 in the process of neurodegeneration. (A) α-Synuclein is taken up into neurons and forms intracellular aggregations with FABP3 and FABP5 [104,105,107,129]. (B) Inflammatory responses induced by exposure to MPP+, rotenone, or transient cerebral ischemia cause α-synuclein and FABP3/5 to accumulate in mitochondria [104,129,132,133]. (C) FABP3 and FABP5 participate in mitochondrial VDAC-1/BAX complex formation to reduce intramitochondrial cytochrome C and membrane potential, resulting in apoptosis [130,133]. ΔΨm: mitochondrial membrane potential; VDAC-1: voltage-dependent anion-selective channel 1; BAX: Bcl-2-associated X protein; mtDNA: mitochondrial DNA.
Figure 3. Schematic diagram of the involvement of FABP3 and FABP5 in mitochondrial dysfunction in neuronal cells of Lewy body disease model. Neurons express FABP3 and FABP5 in the process of neurodegeneration. (A) α-Synuclein is taken up into neurons and forms intracellular aggregations with FABP3 and FABP5 [104,105,107,129]. (B) Inflammatory responses induced by exposure to MPP+, rotenone, or transient cerebral ischemia cause α-synuclein and FABP3/5 to accumulate in mitochondria [104,129,132,133]. (C) FABP3 and FABP5 participate in mitochondrial VDAC-1/BAX complex formation to reduce intramitochondrial cytochrome C and membrane potential, resulting in apoptosis [130,133]. ΔΨm: mitochondrial membrane potential; VDAC-1: voltage-dependent anion-selective channel 1; BAX: Bcl-2-associated X protein; mtDNA: mitochondrial DNA.
Ijms 24 17037 g003
Figure 4. Technique for quantifying risk scores for neurodegenerative diseases using plasma biomarkers. (A) Flow chart from blood collection to plasma preparation and ELISA analysis. In addition to blood collection at the hospital, patients or healthy ones can collect their own blood at home. Plasma samples are analyzed using high-sensitivity immunoassays [31]. (B) Representative quantified data of risk scores for mild cognitive impairment (MCI), Alzheimer’s disease (AD), Parkinson’s disease (PD), and dementia with Lewy bodies (DLB) [31]. Student’s t-test is used to compare the means between each group. **** p < 0.0001 versus healthy controls (HC).
Figure 4. Technique for quantifying risk scores for neurodegenerative diseases using plasma biomarkers. (A) Flow chart from blood collection to plasma preparation and ELISA analysis. In addition to blood collection at the hospital, patients or healthy ones can collect their own blood at home. Plasma samples are analyzed using high-sensitivity immunoassays [31]. (B) Representative quantified data of risk scores for mild cognitive impairment (MCI), Alzheimer’s disease (AD), Parkinson’s disease (PD), and dementia with Lewy bodies (DLB) [31]. Student’s t-test is used to compare the means between each group. **** p < 0.0001 versus healthy controls (HC).
Ijms 24 17037 g004
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

Kawahata, I.; Fukunaga, K. Pathogenic Impact of Fatty Acid-Binding Proteins in Parkinson’s Disease—Potential Biomarkers and Therapeutic Targets. Int. J. Mol. Sci. 2023, 24, 17037. https://doi.org/10.3390/ijms242317037

AMA Style

Kawahata I, Fukunaga K. Pathogenic Impact of Fatty Acid-Binding Proteins in Parkinson’s Disease—Potential Biomarkers and Therapeutic Targets. International Journal of Molecular Sciences. 2023; 24(23):17037. https://doi.org/10.3390/ijms242317037

Chicago/Turabian Style

Kawahata, Ichiro, and Kohji Fukunaga. 2023. "Pathogenic Impact of Fatty Acid-Binding Proteins in Parkinson’s Disease—Potential Biomarkers and Therapeutic Targets" International Journal of Molecular Sciences 24, no. 23: 17037. https://doi.org/10.3390/ijms242317037

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