Next Article in Journal
Evidence of Oxytosis/Ferroptosis in Niemann–Pick Disease Type C
Previous Article in Journal
Association Between Tissue Accumulation of Skin Autofluorescence, Disease, and Exercise Capacity in Older Adults
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phytochemicals Modulate Biosynthesis and Function of Serotonin, Dopamine, and Norepinephrine for Treatment of Monoamine Neurotransmission-Related Psychiatric Diseases

Department of Health and Nutritional Sciences, Faculty of Health Sciences, Aichi Gakuin University, 12 Araike, Iwasaki-cho, Nisshin 320-195, Aichi, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(7), 2916; https://doi.org/10.3390/ijms26072916
Submission received: 12 January 2025 / Revised: 17 March 2025 / Accepted: 21 March 2025 / Published: 23 March 2025
(This article belongs to the Section Bioactives and Nutraceuticals)

Abstract

:
Serotonin (5-HT), dopamine (DA), and norepinephrine (NE) are key monoamine neurotransmitters regulating behaviors, mood, and cognition. 5-HT affects early brain development, and its dysfunction induces brain vulnerability to stress, raising the risk of depression, anxiety, and autism in adulthood. These neurotransmitters are synthesized from tryptophan and tyrosine via hydroxylation and decarboxylation, and are metabolized by monoamine oxidase (MAO). This review aims to summarize the current findings on the role of dietary phytochemicals in modulating monoamine neurotransmitter biosynthesis, metabolism, and function, with an emphasis on their potential therapeutic applications in neuropsychiatric disorders. Phytochemicals exert antioxidant, neurotrophic, and neurohormonal activities, regulate gene expression, and induce epigenetic modifications. Phytoestrogens activate the estrogen receptors or estrogen-responsive elements of the promoter of target genes, enhance transcription of tryptophan hydroxylase and tyrosine hydroxylase, while inhibiting that of MAO. These compounds also influence the interaction between genetic and environmental factors, potentially reversing dysregulated neurotransmission and the brain architecture associated with neuropsychiatric conditions. Despite promising preclinical findings, clinical applications of phytochemicals remain challenging. Advances in nanotechnology and targeted delivery systems offer potential solutions to enhance clinical efficacy. This review discusses mechanisms, challenges, and strategies, underscoring the need for further research to advance phytochemical-based interventions for neuropsychiatric diseases.

1. Introduction

Aromatic amino acids in the brain serve as precursors for the monoamine neurotransmitters, with tryptophan giving rise to serotonin (5-hydroxytryptamine, 5-HT) and tyrosine serving as the substrate for catecholamines, including dopamine (DA), norepinephrine (NE), and epinephrine (E). Monoamine neurotransmitters include indoleamines and catecholamines, and trace amines. They not only transmit neuronal signals, but also have pleiotropic activity, regulate neuronal function, structure, and are implicated in the pathogenesis of neuropsychiatric disorders. As the predominant neurotransmitter among indoleamines, 5-HT regulates brain development via neuronal differentiation, proliferation, synaptogenesis, dendrite organization, and neurogenesis, and it affects mood, anxiety, stress, sexual behaviors, and cognition [1,2]. 5-HT dysfunction during early brain development impairs the synaptic network and impacts the outcome of schizophrenic psychoses, the autism spectrum, and attention-deficit/hyperactivity disorder (ADHD) [3]. The catecholamines DA and NE play key roles in motor function, spatial memory, motivation, arousal, reward, and pleasure. Another catecholamine, E, has a more limited role in central nervous system (CNS) function but is primarily involved in peripheral sympathetic responses. Malfunctions of DA signaling are implicated in the pathogenesis of Parkinson’s disease (PD), Huntington’s disease (HD), ADHD, and schizophrenia [4]. Deficits in monoamines have been targeted for the treatment of depression and PD and beneficial results have been presented.
Monoamine neurotransmission is a complex process, including monoamine biosynthesis, storage, release, interaction with its specific post- and pre-synaptic receptors, uptake, and catabolism [5]. In the brain, indolamines and catecholamines are synthesized and metabolized by similar enzymatic mechanisms. 5-HT and DA are synthesized from L-tryptophan and L-tyrosine, in serotonergic or catecholaminergic neurons, respectively, by two-step reactions: in serotonergic neurons, tryptophan hydroxylase (TPH) catalyzes the hydroxylation of L-tryptophan, followed by decarboxylation by aromatic amino acid decarboxylase (AADC) to produce 5-HT; in catecholaminergic neurons, tyrosine hydroxylase (TH) catalyzes the hydroxylation of L-tyrosine to L-DOPA, which is subsequently decarboxylated by AADC to form DA [6,7]. NE is then synthesized from DA by dopamine-β-hydroxylase (DBH), and E is synthesized from NE by phenylethanolamine-N-methyltransferase (PNMT). Monoamines are metabolized mainly by monoamine oxidase (MAO) and catechol O-methyltransferase (COMT).
Multiple genetic and environmental factors modulate the expression and activity of monoamine-related enzymes and strictly regulate monoamine levels in the brain. TPH and TH activity are regulated by phosphorylation by protein kinases, and feedback inhibition by the monoamines and allosteric regulation [8]. Stress influences TPH and TH expression at multiple stages [9]. Expression of neuron-specific TPH2 is mainly regulated by stress, cortisol, and 5′- and 3′-regulatory polymorphisms of TPH2 [10]. TH and TPH expression and activity are regulated by estrogen and glucocorticoids [11]. Diet, vitamins, and nutraceuticals modulate monoamine biosynthesis and the metabolism of 5-HT in the brain. The 5-HT biosynthesis rate mainly depends on the brain’s L-tryptophan levels [7,12]. AADC requires pyridoxal phosphate, DBH ascorbic acid, and PNMT S-adenosylmethionine (SAM) as a methyl donor for enzymatic activity.
Diet affects age-related neurodegeneration and cognitive decline, and ingredients of plant food have been proposed as neuroprotective molecules [13]. Phytochemicals are secondary plant metabolites and polyphenols are a major member, including flavonoids and several classes of non-flavonoids such as phenolic acids, stilbenes, and lignans [14]. These compounds exhibit various biological effects, such as antioxidant activity, regulation of cellular signaling and gene induction, neuroprotection, immune-regulation, and modulation of gut microbiota. Neuroprotective polyphenols modulate neurotransmission systems and ameliorate neurochemical and behavioral changes associated with neuropsychiatric disorders [15,16,17]. Some bioactive polyphenols, known as phytoestrogens, have biological activity similar to 17β-estradiol (E2). They bind to the estrogen receptors (ERs) and exert pro- and anti-estrogenic effects [18]. They modulate monoamine transmitter systems via affecting the expression and activity of related enzymes, TPH, TH, AADC, and MAO [19].
This paper presents the therapeutic potential of phytochemicals against neurobehavioral disorders caused by deregulated homeostasis of monoamine neurotransmission. The genetic and environmental factors regulate monoamine biosynthesis, metabolism, and function, and their deregulation in the neuropsychiatric disorders is discussed [1,20]. Multiple neuroprotective functions of polyphenols, such as antioxidant, anti-apoptosis functions, induction of pro-survival genes, and neurotrophic factor (NTF)-like activity are reviewed [21,22,23,24,25]. The epigenetic effects of phytochemicals on the interaction between genetic and environmental factors in the early phase of brain development is discussed. A deficit or excess of 5-HT levels in the developing brain is associated with depression, aggressive behaviors, and anxiety in adulthood. The strategy to apply phytochemicals for the prevention and therapy of neuropsychiatric diseases is discussed.

2. Biosynthesis, Metabolism, and Function of Monoamine Neurotransmitters and Related Enzymes in the Brain

5-HT is synthesized in only a small number of neurons localized in the midbrain dorsal and ventral raphe nuclei, but the 5-HT release site and synapses are distributed throughout the brain. TPH hydroxylates L-tryptophan into 5-hydroxytryptophan (5-HTP) using Fe2+, (6R)-L-tetrahydrobiopterin (BH4), and diatomic oxygen (O2) (Figure 1). 5-HTP is decarboxylated into 5-HT by AADC. 5-HT is metabolized by type A MAO (MAO-A) to 5-hydroxyindol-acetaldehyde (5-HIAL), which aldehyde dehydrogenase (ALDH) converts to 5-hydroxyindolacetic acid (5-HIAA). The 5-HT system is implicated in regulating human behaviors, state of mind, mood, social interaction, and consciousness, and is involved in the outcome of schizophrenic psychoses, the autism spectrum, and ADHD [3].
TH and TPH are members of aromatic amino acid hydroxylase (AAAH) family monooxygenases and perform the hydroxylation of the aromatic ring of an amino acid by use of diatomic oxygen, a catalytic non-heme ferrous iron, and BH4 as a coenzyme [26]. They have very similar active sites and are composed of a multi-domain structure: the N-terminal regulator domain (R) of 100–150 amino acid residues, the catalytic domain (C) of about 330 residues, and a coiled-coil domain at the C-terminal of about 20 amino acids. There are two TPH isoenzymes, TPH1 and TPH2, encoded in different genes localized on chromosomes 11 and 12, respectively, with sequence identity of 71% [27]. TPH2 is expressed predominantly in the brain, whereas TPH1 in peripheral tissues. TPH2 contains a larger regulatory domain than TPH1, and an additional 41 amino acids at the N-terminus. A serine at position 16 (Ser16) of TPH2 is phosphorylated by cAMP-dependent protein kinase A (PKA) and increases the activity, interacts with 14-3-3 proteins, and increases protein stability [28]. There are more than 300 single-nucleotide polymorphisms (SNPs) of TPH that have been identified in humans, six are coding non-synonymous SNPs (L36P, P206S, A328V, R441H, D479E), and three are coding synonymous SNPs (P312P, L327L. A355). Some of these polymorphisms are associated with major depressive disorder (MDD), suicide, and obsessive-compulsive disorder [29,30], but other studies presented lack these associations [31]. 5-HT synthesis in the brain is highly limited by serum L-tryptophan levels.
The DA neurons are mainly localized in the substantia nigra (SN) pars compacta, ventral tegmental area of the midbrain, and the hypothalamus. DA biosynthesis is regulated by the activity of TH and guanidine triphosphate cyclohydrolase I (GTPCH), the rate-limiting enzyme of BH4 biosynthesis. MAO-A and -B oxidize DA into 3,4-dihydroxyphenlacetaldehyde (DOPAL), which is oxidized by ALDH to 3,4-dihydroxyphenylacetic acid (DOPAC), or reduced by aldehyde reductase (ARL) to 3,4-htdroxyphenylglycol (DHPG). COMT converts DA into 3-methoxytyramine (3-MT) and NE, E and L-DOPA into normetanephrine (NMN), and metanephrine (MN) and 3-methoxydopa (3-MD), respectively (Figure 2).
The human TH (hTH) gene is localized on chromosome 11p15.5 and has 14 exons. The hTH is transcripted to alternative splicing at the 5′ end and produces four isoforms of hTH 1, 2, 3, 4 with different R domains [32]. TH activity is regulated by feedback inhibition via allosteric modulation by polyanions and posttranslational phosphorylation and dephosphorylation. DA, NE, and E are feedback inhibitors of TH via binding to the active site though the iron atom. Phosphorylation of Ser40 in the R domain by PKA increases the dissociation of catecholamines from the iron by 2–3 orders of magnitude, and decreases the Km value for BH4 by 2-fold [33]. The phosphatidylinositol-3 kinase (PI3K) and ERK1/2 pathways increased phosphorylation of Ser31 of TH, protein kinase C (PKC) Ser19, and PKA Ser40, respectively. Ser40 phosphorylation increased TH activity [34].
The TH protein is found to decrease markedly in the SN of Parkinsonian patients. TH mutations have been speculated to modify the susceptibility of the sporadic form of PD [35]. Three SNPs (rs2072056, rs6356, rs10743152) were reported within the TH gene, but the association with PD risk was not fully proven [32]. hTH deficiency is an autosomal recessive disorder due to mutations in the TH gene, and fewer than 40 patients were reported worldwide. According to neurological features, two phenotypes of TH deficiency were reported. Type A presents progressive hypokinetic-rigid syndrome with dystonia of infantile onset, whereas type B presents a complex encephalopathy of neonatal onset [36]. TH expression is induced by acetylcholine through nicotinic cholinergic receptors, and by vasoactive intestinal polypeptide (VIP) and pituitary adenylate cyclase-activating peptide (PACAP) in PC12 cells and sympathetic ganglia. The TH gene promoter contains cAMP responsive elements (CREs) and activates the protein-1 (AP-1) site [37], and TH transcription is regulated by PKA signal transduction [38].
The AADC catalyzes the decarboxylation of L-DOPA, 5-hydrotryptophan, and, much less efficiently, aromatic amino acids (tyrosine, tryptophan, phenylalanine) and histidine [39]. The AADC gene is localized in the catecholaminergic and serotonergic neurons of the central and peripheral nervous system and in chromaffin cells of the adrenal medulla. The human AADC is localized at 7p12.2-p.12.1 and has 15 exons and 14 introns [40]. Missense variants, frameshift variants, and alterations of splicing sites cause AADC deficiency. AADC deficiency is a rare congenital autosomal disorder, decreases 5-HT and DA biosynthesis, and causes neurodevelopment delay, hypotonia, oculogyric crises, and movement disorders [41].
MAO is classified into type A and B (MAO-A, MAO-B), according to substrate specificity and inhibitor sensitivity. MAO-A has a higher affinity to 5-HT, NE, and E, and MAO-B to phenylethylamine (PEA), benzylamine (BA), and octopamine. DA, tyramine, and tryptamine are common substrates for both MAO-A and -B. MAO-A is localized in catecholaminergic neurons and MAO-B in serotonergic and histaminergic neurons. Deregulated MAO-A activity is implicated in the pathogenesis of depression and antisocial behaviors, suggesting the possible application of a MAO-A inhibitor as a tool for treatment [42]. MAO-B increases with age and promotes neurotoxic reactive oxygen species (ROS), which is one of the pathogenic factors in PD and Alzheimer’s disease (AD). MAO-A regulates the 5-HT levels in an embryonic brain and modulates the development of brain architecture [43].
Human MAO-A and -B are coded by distinct genes localized on the X-chromosome (Xp11.23) and the MAO-A gene contains the MAO-A gene-linked polymorphic region (MAO-A-LPR) positioned about 1.2 kb upstream of the MAO-A transcription initiation site. The MAO-A variable number tandem repeat (MAO-A-VNTR) polymorphism codes 3, 3.5, 4, or 5 copies of the 30 bp repetitive sequence and affects the transcriptional activity of the MAO-A gene promoter [44]. Alleles with 3.5 or 4 copies of the repeated sequence (MAO-A-H) are transcribed 2–10 times more efficiently than 2, 3, or 5 copies (MAO-A-L). The MAO-A polymorphism drives variations in MAO-A activity and influences impulsivity and aggression. The MAO-A-L enhances the risk for developing an aggressive, antisocial personality. Low MAO-A expression causes 5-HT excess and impairs critical neural circuitry for social evaluation and emotion regulation, resulting in amplifying the effects of adverse early life experiments [45]. MAO-A-H carriers are at a greater risk of depression [46], but the results are still conflicting. A MAO-A deficit in Brunner syndrome caused by a nonsense mutation of the MAO-A gene (rs72554632) and Norrie disease causes behavioral abnormality, such as episodic impulsive aggression and borderline mental retardation [47,48].
COMT catalyzes the methylation of catechol substances, including catecholamines, catechol estrogens, and polyphenols, and is expressed throughout the brain. COMT exists in two forms: a membrane-bound form in the brain and a soluble form in the peripheral tissues. A membrane-bound COMT has higher affinity to DA in the human prefrontal neurons [49]. COMT plays a role in cognition, behavior, emotion, pain processing, addictive behavior, and neurodegeneration, and is implicated in neuropsychiatric disorders in a sex-different way [50,51]. The COMT gene is localized in chromosome 22q11, one of the principal loci linked to schizophrenia. Val158Met functional polymorphism plays a primary role in DA metabolism in the prefrontal cortex, and is associated with negative symptoms in patients with schizophrenia and bipolar disorder [52].

3. Neuroprotective Effects of Phytochemicals in Brain Health

Phytochemicals have multiple neuroprotective activities, such as antioxidant, anti-inflammatory, antiapoptotic activity, mitochondrial stabilization, activation of cellar signaling pathway, and induction of neuroprotective genes [22,53]. Some of them can cross the blood–brain barrier (BBB) and show protective function in neurodegenerative disorders [54]. The phytochemical superfamily consists of polyphenols (flavonoids, non-flavonoids), terpenoids/isoprenoids (saponins, lycopene, etc.) and nitrogen-containing alkaloids (caffeine, morphine, nicotine, etc.). Bioactive polyphenols are major phytochemicals composed of multi hydroxyl (-OH) aromatic phenols. Approximately 8000 polyphenols have been identified, and more than 4000 belong to flavonoids [55]. Flavonoids are subclassified into flavonols (quercetin, kaempferol, myricetin, etc.); flavones (luteolin, apigenin, etc.); isoflavones (daidzein, genistein, etc.); flavanones (naringenin, hesperetin, etc.), flavanols [catechin, epicatechin (EC), gallocatechin (GC), epigallocatechin (EGC), its gallate (EGCG), etc.], chalcone; and anthocyanidins (malvidin, cyanidin, etc.). Non-flavonoids include stilbene (resveratrol), lignan (enterolactone), tannin (Proanthocyanidin A1), and phenolic acids (ferulic, ellagic, tannic, gallic and caffeic acids, etc.) [14] (Figure 3).
Oxidative stress and inflammation are common features of neurodegenerative disorders and aging. Polyphenols have broad antioxidant activity; scavenging free radicals, metal chelating, mitochondrial protection, and suppression of ROS-generating enzymes. The structure required for flavonoids and phenolic acids to scavenge radicals and chelate metals include the catechol group (3′, 4′-hydroxy groups in ring B, C2-C3 double bond in conjugation with a C4-keto function in ring C and presence of 3- and 5-hydroxyl groups in ring C and A) [56,57] (Figure 4). Flavonoids, curcumin, and caffeic acid alkyl esters regulate the gene expression of antioxidant enzymes [superoxide dismutase (SOD), catalase, glutathione reductase (GR), glutathione peroxidase (GPx)] by nuclear erythroid 2-related factor 2/antioxidant response element (Nrf2/ARE) signaling pathways [58,59]. Flavonols activate Nrf2, a transcription factor, and increase ARE-regulated genes.
Polyphenols activate PI3K, Akt/protein kinase B (Akt/PKB), tyrosine kinases, PKC and mitogen activated protein kinase (MAPK) signal pathways, and affect cellular function by modulating gene expression and phosphorylating targeted molecules [60,61]. Flavonoids selectively interact with MAPK signaling pathways to activate the expression of neuroprotective genes involved in NTF-induced differentiation, apoptosis, and various forms of cellular plasticity. Stress signals (DA, 4-hydroxy-2-nonenal, ROS, inflammatory kinases) activate c-Jun-N-terminal kinase (JNK) and p38, leading to apoptosis. Flavonols, epicatechin, and baicalein, suppress JNK and downstream c-jun and pro-caspase-3, and protect neurons.
Polyphenols exert NTF-like actions through direct binding to NTF-receptors, activating cellular signaling pathways, and inducing NTF expression [62,63]. 7,8-Dihydroxyflavone has a direct agonistic effect on the tropomyosin receptor kinase (Trk) receptor for the brain-derived neurotrophic factor (BDNF) and nerve growth factor (NGF). Flavonoids activate ERK and PI3K/Akt signaling pathways and increase NTF expression, whereas apigenin, ferulic acid, and resveratrol increase cAMP response element-binding protein (CREB) phosphorylation, and increase BDNF and glial cell line-derived neurotrophic factor (GDNF). High flavonoid intake induced serum BDNF levels, improved cognitive function in healthy subjects [64], and, in women with premenstrual syndrome [65] in randomized, double-blind, placebo-control trials, curcumin increased serum BDNF levels.
Polyphenols have been proposed to maintain mitochondrial homeostasis and ATP synthesis, modulate apoptosis systems in mitochondria, and protect neurons [23,66]. Resveratrol activated sirtuin-1 (SIRT1), decreased acetylation of peroxisome proliferator-activated receptor γ coactivator α (PGC-1α), increased genes for oxidative phosphorylation and mitochondrial biogenesis, and improved mitochondrial functions [67]. Mitochondria are directly involved in neuronal programmed cell death. Apoptosis progresses sequentially via the opening of the mitochondrial permeability transition pore (mPTP), resulting in the release of calcium and apoptogenic proteins [cytochrome c (Cytc), Smac/DIABLO, Omi] into cytosol, activation of caspases and subsequent chromatin condensation, and DNA fragmentation [68]. The mPTP is a complex apparatus mainly composed of the voltage-dependent anion channel (VDAC), the outer membrane transporter protein (TSPO), the Bcl-2 protein family at the outer mitochondrial membrane (OMM), the adenine nucleotide translocator (ANT) at the inner mitochondrial membrane (IMM), and cyclophilin D (CypD) at the matrix. By using the cellular model of apoptosis induced by a TSPO ligand, PK11195, the process of the mPTP opening was followed [69,70]. PK11195 opens a transitionally reversible pore at the IMM, declines mitochondrial membrane potential (ΔΨm), and allows the influx of water and molecules of less than 980 Da. Excess stimulus irreversibly forms the mPTP and releases molecules up to 1500 Da, such as Cytc. Phytochemicals such as ferulic acid, its derivatives, astaxanthin, and some flavonoids, prevent the pore formation of ANT with CypD at the IMM, thereby inhibiting mPTP formation and apoptosis [24]. Astaxanthin and resveratrol decreased the expression of CypD and ANT in heart mitochondria isolated from rats treated with isoproterenol [71] and after ischemia-reperfusion [72]. Resveratrol downregulated VDAC expression, dephosphorylated or deacetylated VDAC, prevented the mPTP opening induced by myocardial ischemia-reperfusion injury [73,74]. Curcumin binds to and stabilizes VDAC in the closed state, influencing mitochondrial function, apoptosis, and mPTP regulation [75]. However, quercetin has ambivalent redox activity and can inhibit or induce the mPTP [76]. (In the Supporting Information, Table S1 presents the chemical compounds and their function on affecting mPTP formation).

4. Sex- and Stress-Related Hormones Modulate Monoamine Neurotransmission in the Brain

The monoamine neurotransmitter system responds to extrinsic and intrinsic stimuli, and adapts its function and expression to maintain brain function [77]. Monoamine neurotransmission is regulated by sex-related estrogens and androgens (testosterone, dihydrotestosterone), and stress-induced corticoids, neurotransmitters, NTFs, ROS, and calcium [11]. Estrogens impact the cellular differentiation and neural network formation, including 5-HT and DA systems. E2 directly modulated promoter activity of TPH2, 5-HT transporter (SERT, SLC6A4), TH, DBH, and GTPCH and increased expression [78,79]. Sex chromosomes (XX versus XY) and sex hormones (estrogen versus testosterone) are the factors for sex differences in neuropsychiatric disorders [80,81,82]. Stress activates the hypothalamic–pituitary–adrenal (HPA) axis and induces TPH2 expression by cortisol released from stress-activated adrenal secretions, which enhances 5-HT and inhibits adrenal cortisol secretion by ACTH [10].
Estrogens are mainly produced in the ovaries, the placenta in females, and the adrenal cortex of males, and are retained in target cells by estrogen receptors (ERs) [83]. Four estrogens, estrone (E1), 17β-estradiol (E2), estriol (E3), and estetrol (E4), have been identified in humans. There are three types of ERs, classical ERα, ERβ, and non-classical G protein-coupled estrogen receptor 1 (GPER1). ERα and ERβ are encoded by ESR1 on chromosome 5 and ESR2 on chromosome 14, respectively. ERα and ERβ exhibit a high degree of homology. ERβ is expressed in the brain, cardiovascular system, lung, and immune system, whereas ERα in the uterus, ovary, bone, white adipose tissue, and liver.
Estrogens bind and change ER conformation, allowing it to interact with specific estrogen response elements (EREs) on the promoter of the target gene, and modulate gene transcription. In classical pathways, ERs form homo- or hetero-dimers, and directly interact with ERE [84,85]. ERs interact with transcription factors, such as AP-1, simian virus 40 promoter factor 1 (Sp1), CCAAT/enhancer binding protein β (C/EBPβ), and nuclear factor κB (NF-κB), and indirectly activate transcription [86,87]. E2 activates tyrosine kinase (SrcK)/MAPK and PI3K/AKT, signaling pathways, phosphorylates, and activates certain nuclear transcription factors.
Interaction between 5-HT and estrogen is indicated by the colocalization of both systems in the same brain regions and is involved in the pathophysiology of depression, migraines, irritable bowel syndrome, eating disorders, and pregnancy-related pathologies [88,89]. Estrogen activates the nuclear and membrane ERβ and induces TPH2 expression through an ERE half-site located within the TPH2 promoter’s 5′-untranslated regions [90]. Estrogen and progesterone increased TPH2 mRNA in the dorsal raphe region of female macaques [91], and estrogen increased TPH2 in the rat raphe nuclei and ameliorated anxiety-like behavior [92]. In female and male ERβ -knockout (KO) mice, 5-HT levels decreased significantly in the hippocampus and nucleus accumbens, and mice showed anxiety-like behavior [93,94]. Estrogens exert antidepressant activity via multiple mechanisms: the regulation of 5-HT synthesis, metabolism and function, and neurotrophic activity to promote neuroplasticity and neurogenesis [95]. Estrogen increases the expression of TPH2, TH, and SLC6A4, downregulates MAO and the 5-HT1A receptor and presents antidepressant-like effects in female patients with depression [79]. E2 increased 5-HT2A and SERT in the thalamus and hypothalamic nucleus of female ovariectomized rats [96]. E2 increased the 5-HT transporter in the superior frontal cortex, anterior cingulate cortex, nucleus accumbens, and 5-HT2A receptor in the frontal cortex and striatum of ovariectomized Macaca fascicularis [97,98].
Estrogen influences neural network formation in the SN and ventral tegmental. The TH promoter contains AP-1/early growth response gene 1 (Erg) motifs, which are required for E2-mediated TH induction by ERβ, but not ERα [99,100,101]. Estradiol increased TH and DBH levels in the locus coeruleus and nucleus solitarius, but not in the SN of ovariectomized female and castrated male rats [101,102,103,104]. E2 increased TH promoter activity and expression in the locus coeruleus of female mice but decreased them in males [105]. E2 interacted with ERα at the plasma membrane, activated PKA or MAPKs, phosphorylated CREB, modulated CRE-mediated TH transcription in PC 12 cells [106], and increased catecholamine synthesis in cultured bovine adrenal medullary cells [107,108].
Estrogen binds to ERβ, interacts with ERE in the MAO-A promoter and downregulates MAO-A expression. MAO-A is associated with sexual dimorphism in several neuropsychiatric disorders. The high prevalence of depression and dementia in peri- and post-menopausal women suggests the modulation of monoamine metabolism by estradiol [109]. MAO-A, assessed by [11C]-harmine (a β-carboline MAO-A inhibitor) position computed tomography (PET) imaging, increased by 43% in the brain of postpartum women, whose estrogen levels dropped by 100–1000-fold during the first 3 to 4 days [110]. In perimenopausal women, MAO-A increased by 34% and 16% compared with reproductive and menopausal women, respectively [111]. Estrogen significantly decreased MAO-A activity in the hypothalamus (−28%) and amygdala (−21%) of adult female ovariectomized rats, whereas MAO-B activity in the brain did not change [112]. ERβ regulates MAO-A expression in the dorsal raphe and paraventricular nucleus, ERα regulates MAO-B in the preoptic area, and both ERα and ERβ regulate MAO-A and MAO-B expression in the ventromedial nucleus [113]. The MAO-A promoter consists of the Sp1 binding site and sex-determining region Y (SRY) binding site [114]. Sp1 and Sp4 directly interact with Sp1 sites and activate the MAO-A and MAO-B promoters, whereas Sp3 inhibits them by competition for binding to the CACCC element [115] (Figure 5). Androgen and glucocorticoid upregulate MAO-A expression, through binding to the androgen receptor (AR) or glucocorticoid receptor (GR), and direct interaction with the functional androgen/glucocorticoid response element (ARE/GRE) [116]. High-dose testosterone treatment decreases MAO-A in human brains [117]. E2, estrogen metabolite, and 4-hydroxeuilenin (4-OHEN) downregulate COMT expression and are substrate sand inhibitors of COMT in MCF-2 cells [118,119,120].
Exposure to stressors causes an increased release of monoamine neurotransmitters in the limbic structures and activation of the HPA-axis, leading to an increase in circulating glucocorticoids. Glucocorticoids bind and activate GRs, and subsequently interact with GRE on the targeted gene. TPH2 exhibits flexible gene expression in response to stressful events or stressors [10]. The 5′-untranslated region of TPH2 and TH contains a binding motif, the repressor element-1 (RE1) for the repressor element-1 silencing transcription factor (REST). This motif functions as an “on-off” switch for THP2 and TH transcription [121]. Glucocorticoids modify TPH2 expression and 5-HT synthesis in a species-specific way. In the raphe nuclei of ovariectomized female and intact male mice, dexamethasone decreased TPH expression [122], whereas in rats, glucocorticoids induced TPH expression and TH synthesis [123]. A daily fluctuation of glucocorticoids regulates circadian rhythmic TPH expression in the raphe nucleus. Stress induces TH and DBH expression in the sympathetic ganglia and adrenal medulla, depending on stress type, animal species, and target tissues in animal models of stress [124]. A glucocorticoid response element was reported at AP-1-like sequences on the rat TH gene [125]. Chronic stress-induced glucocorticoids increased Kruppel-like factor 11 [KLF11, also called transforming growth factor β-inducible early gene (TIEG2)], translocated KLF11 from the cytoplasm to nucleus, and activated Sp/KLF-binding sites at the MAO-A promoter and increased the expression of MAO-A in the rat brain [126].

5. Phytoestrogens Promote Expression and Activity of Tryptophan Hydroxylase and Tyrosine Hydroxylase

The discovery of ERβ has provoked the search for ER type-specific ER modulators (SERM) for the prevention or treatment of menopausal symptoms, osteoporosis, breast cancer in women, and other estrogen-related disorders. ERβ-selective SERM (coumestrol, diarylpropionitrile) administration in the hippocampus of ovariectomized rats decreased anxiety and depressive behaviors [127]. However, ERβ-targeted SERM has not been available in practice, and applications of plant bioactive compounds have been searched for to prevent neuropsychiatric disorders, cardiovascular disease, cancer, and metabolic syndrome [128]. Phytoestrogens are nonsteroidal estrogenic polyphenols derived from dietary plants. They include isoflavones (daidzein, genistein, biochanin A), flavonoids (chrysin, apigenin, apigenin, naringenin, kaempferol, quercetin), stilbenes (trans-resveratrol), coumestrol, and lignans. Phytoestrogens have been proposed as part of hormone replacement therapy in estrogen-related disorders [129]. Soybean-derived isoflavones have some benefits for menopausal syndromes, such as vasomotor syndromes and hot flashes, but clinical trials could not confirm the effects on menopausal syndromes [130,131].
Phytoestrogens bind ERs at the membrane and nucleus, induce estrogen-dependent gene transcription, and have biological activity similar to E2 [18]. Some of them (coumestrol, genistein, apigenin, naringenin, kaempferol) have a higher binding affinity to ERβ than ERα and stimulate transcriptional activity [132,133]. Among flavonoids, genistein (4′,5,7-trihydroxyisoflavone) exerts the most potent estrogenic activity to ERβ, followed by daidzein (4′,7-dihdroxyisoflavone), biochanin A (4′,5-dihdroxy-7-methoxyisoflavone), apigenin (4′,7-dihdroxyflavone), kaempferol, naringenin, quercetin, and chrysin. The position and number of the hydroxyl substituents on the flavone or isoflavone molecule determine the ER binding affinity. 8-Prenylnaringenin, biochanin A, daidzein, genistein, naringenin, resveratrol, and quercetin modulate plasma estrogen levels in pre- and post-menopausal women and male/female volunteers [134]. (In the Supporting Information, Figure S1 presents the chemical structure of estrogenic phytochemicals).
Phytoestrogens modulate the serotonergic and dopaminergic systems, and improve brain function and neuropsychiatric diseases [15]. They affect the activity and expression of TPH, TH, and MAO and regulate the monoamine biosynthesis, metabolism, and function. In addition, phytochemicals control the transporters and receptors of monoamine neurotransmitters. In an aged brain, oxidative stress and inflammation decreased TPH and TH phosphorylation and their activities [135,136]. Silymarin, quercetin, naringenin [137], catechin, and tea extract (polyphenon-60, a catechin extract) [138] enhanced the activity of TPH and TH in the brain and adrenal medullary cells. Chronic resveratrol treatment (20 mg/kg/day for 4 weeks) in old male rats (20 months) increased the activity of TPH2 (70–51%) and TH (150–36%) in the hippocampus and striatum; TPH1 activity (463%) in the pineal gland; enhanced 5-HT, NE, and DA levels; and improved cognitive and motor functions [139]. Korean red ginseng increased TPH2 and TPH1 in the hippocampus of a rat model of prolonged stress and ameliorated depression-like behavior [140,141,142]. Anthocyanin extract from blueberries increased the TPH protein in the hippocampus of aged male rats [143].
Puerarin extracted from Pueraria lobata increased TH expression in the SN by about 85% in rats [144]. Sesamol and naringenin increased TH expression in the SN of a rotenone-treated PD rat model [145]. Chlorogenic acid, a trans-cinnamic acid ester, increased TH expression in the SN and striatum of MPTP-treated mice [146]. Resveratrol increased TH protein expression in the striatum of pups exposed to lipopolysaccharide in utero from dams fed with a resveratrol-supplemented diet [147]. Phloretin (dihydronaringenin) found in apples increased TH protein expression in the striatum of MPTP-treated mice and protected DA neurons by anti-inflammatory activity [148]. Tangeritin, a citrus polymethoxy flavone, increased TH expression in a Drosophila model of PD [149]. Curcumin increased expression of TH mRNA, DA, and NE levels and downregulated MAO in the limbic system and midbrain of ovariectomized rats [150]. The non-planar structure and hydroxy groups on aromatic rings of curcumin seems to promote TH transcription. Gardenin A, a polymethoxy flavone isolated from Cardenia resinfera, regulated the nuclear factor erythroid 3-related factor (Nrf2), and increased TH protein expression in the striatum of A53T α-synuclein-overexpressing mice [151].
Daidzein and resveratrol activated TH activity in cultured adrenal medullary cells through ERK1/2 activation [19]. Daidzein at low concentrations enhanced catecholamine neurotransmitter synthesis via binding to plasma membrane ERs, but at high concentrations, inhibited the biosynthesis [152]. Nobiletin, a citrus polymethoxy flavone, phosphorylated Ser19 and Ser40 of TH and increased TH activity in cultured bovine adrenal medullary cells by PKA [153]. (In the Supporting Information, Table S2 presents the chemical structure and function of phytochemicals affecting TPH and TH).

6. Phytochemicals Inhibit Expression and Activity of Monoamine Oxidase and Catechol-O-Methyltransferase

MAO plays crucial physiological roles and MAO-A is a target for the therapy of depression and anxiety, whereas MAO-B for that of PD and AD. Synthetic, irreversible MAO-A inhibitors (clorgyline) showed critical side effects, and reversible MAO-A inhibitors (moclobemide, brofaromine, toloxatone) are clinically applied for the treatment of anxiety and depression. Irreversible MAO-B inhibitors, selegiline [(−)-deprenyl] and rasagiline and their derivatives, are used for the treatment of neurodegenerative diseases, including PD, AD, and aging. Herbs and herb preparation inhibit the activity of MAO and COMT, and are considered as effective alternative therapies in neuropsychiatric diseases [154]. Herb ingredients, including flavonoids (flavanols, flavanones, flavones, isoflavones flavonols), alkaloids, chalcones, coumarins, and xanthones (mangiferin), have been proposed to inhibit MAO-A and exert antidepressant-like and neuroprotective functions [155,156].
Quercetin and structurally related flavonoids can cross the BBB, and attenuate MAO-A in the brain to exert antidepressant-like effects in rodent models of depression [157]. Chemical structures for specific MAO-A inhibition by flavonoids are indicated. Hydrophobicity and the planar structure of the diphenyl-propane (C6-C3-C6) skeleton are required for effective inhibition against MAO-A. A flavone with a planar structure inhibits MAO-A better than a flavanone with a nonplanar skeleton [158]. Hydroxy residue at the C-4′ position of the B ring of flavonoids increases the affinity to MAO-A [159]. Quercetin, kaempferol, purpurin, and apigenin are potent, selective, reversible, competitive MAO-A inhibitors [160,161], whereas chrysin is a potent MAO-B inhibitor [162]. Alizarin (1,2-dihydroxyanthracene-9,10-dione) inhibited MAO-B, and less potentially, MAO-A. Resveratrol and (−)-trans-e-viniferin (a resveratrol dimer), natural coumarin, xanthotoxin, and praeruptorin-A have MAO-A and -B inhibition and antidepressant activity [163]. Xanthines (caffeine), flavonoids (gancaonin A), protocatechuic acid, and alkaloids (piperine) are potent reversible, selective MAO-B inhibitors [164]. (Figure S2 presents the chemical structure and function of phytochemicals affecting MAO in the Supporting Information).
Polyphenols are used as scaffolds for the development of new MAO inhibitors. Open chain flavonoids composed of a core structure of 1,3-diaryl-2-prone-1-one are derivatized as MAO-A and -B inhibitors. Derivatives of natural and synthetic chalcone (1,3-diphenyl-2-propaen-1-one) inhibit MAO, whereas those of heterocyclic (furan, thiophene, piperidine, quinoline) are potent, reversible MAO-B inhibitors [165]. 3-Phenylcoumarin derivatives, such as aminopyran, are synthesized to develop new antidepressants [166]. The alkyl-sulfonyl group substitution at the C-7 position of coumarin provided MAO-A inhibiting activity in esuprone and 7-oxycoumarin [167,168,169]. A benzyloxy substitution increased selective MAO-B inhibition in 7-benzyloxy-3,4-dimethlcoumarin and 3-arylcoumarin derivatives [168,170].
Several natural and synthetic COMT inhibitors have been developed to increase the available monoamine neurotransmitters for therapy for PD, depression, and schizophrenia [170]. The derivatives of pyrogallol and catechol, such as gallic acid, caffeic acid, 2-hydroxy estradiols, or flavonoids (quercetin, rutin), are so-called “first-generation” COMT inhibitors. Derivatives of nitrocatechol, chalcone, and pyrazoline have been developed as the “second generation” of COMT inhibitors [171]. Entacapone and opicapone have been synthesized for adjunctive therapy of PD. COMT inhibitors with low toxicity and high bioavailability have been searched among catechol and DA derivatives. Catechol polyphenols, (+)-catechin, alizarin, morin, chlorogenic acid, and fisetin inhibited COMT in rat liver [172]. Chlorogenic acid [173] and catechins with a galloyl-type D-ring, EGCG, and (−)-epicatechin-3-gallate were the most potent human liver COMT inhibitors [174]. (Figure S3 presents the structures of cited compounds that inhibit COMT in the Supporting Information).

7. Epigenetic Phytochemicals Affect Monoamine Neurotransmission and Prevent and Treat Monoamine-Related Disorders

Neurodevelopment is under continuous endogenic and exogenic stress from gestation to adulthood. Life experiences influence behavior, stress response, and disease susceptibility through the epigenetic modification of expression of genes, and subsequently, enzymes and molecules [175]. Nutrients and bioactive food components show epigenetic modification, and an “epigenetic diet” has been proposed to delay the onset of aging and age-related neurodegeneration [176]. The epigenome begins encoding in the uterus and epigenetic mechanisms are associated with neurogenesis, cell migration, and synaptogenesis in early brain development [177]. Periconceptional nutrients can affect the epigenomic state of offspring during the first 1000 days, from conception to the end of age 2, and significantly affect neurodevelopment. The perinatal nutritional status causes permanent changes in insulin-like growth factor 2 (IGF2) methylation and the newborn’s fetal growth [178]. The maternal intake of flavonoid (kaempferol-3-O-glucoside, narirutin) reverted depression-like behavior in female rat offspring [179]. Dietary phytochemicals may modify the epigenome, intervene in the development of monoamine-related disorders, and help prevent their progression.
Polymorphisms of TPH2, SERT, and MAO-A, and exogenous factors, such as stress, physical abuse, nutrition, tryptophan depletion, and poor maternal care, influence 5-HT systems. 5-HT-dependent signaling regulates the early phase of brain development [180]. In the human brain, 5-HT neurons are identified by 5 weeks of gestation ahead of other monoamine systems, and proliferate until gestational week 10 [181] and regulate the development of other neurotransmitter systems [1]. 5-HT levels increase during the first 2 years of age and decline to adult levels by age five. Decreased or increased 5-HT levels affects the structure of brain circuits and is implicated in psychiatric disorders [182,183]. Prenatal treatment with 5,7-dihydroytryptamine decreased 5-HT signaling, induced maladaptive behaviors, including depressive-like and anxiety-like behavior, and defective social interactions in adult rats [20]. Licking/grooming in rats increased 5-HT, activated the 5-HT7 receptor, promoted NGF-inducible factor A (NGFI-A), demethylated the GR promoter, and modulated the HPA response to stress [184]. The depletion of 5-HT decreased neurogenesis, induced abnormal dendritic density in the hippocampus, and impaired spatial learning [185]. Excess 5-HT during the early stage disrupts the normal wiring of the somatosensory cortex. In MAO-A or SERT KO mice, the thalamocortical axons failed to segregate and did not form a normal barrel-like structure [186]. 5-HT excess caused by a MAO-A deficit is involved in antisocial and aggressive behavior, as shown in Brunner syndrome and MAO-A KO mice [20]. In MAO-A KO mice, forebrain-expression of MAO-A reduced 5-HT, NE, and DA levels, restored the brain structure, and rescued aggressive behavior [187].
Epigenetic changes are mediated by DNA methylation, histone modification, non-coding RNA (ncRNA), and a changed structure of chromatin (Figure 6). DNA methylation occurs on cytosines at CpG sites by DNA methyltransferases (DNMTs) using SAM as a methyl donor, and represses gene expression. In mammalian cells, the N-terminal histone tails are post-translationally modified by acetylation, methylation, phosphorylation, and ubiquitination. Histone acetylation is controlled by histone acetyltransferases (HATs) and histone deacetylases (HDACs). HATs add chromatin acetyl groups, make chromatin conformation more relaxed, and promote gene transcription, whereas HDACs act in reverse ways. ncRNAs are classified into microRNAs (miRNAs) and long noncoding RNAs (lncRNAs) by their length. miRNAs are small, single-stranded RNAs with around 22 nucleotides in length, bind to the target mRNA at the 3′-untranslated regions (3′UTRs), degrade, destabilize mRNA, inhibit translation, and silence gene expression. lncRNAs are RNAs with more than 200 nucleotides in length, modulate gene expression at the transcriptional and post-transcriptional levels, and modulate chromatin architecture. lncRNAs are strongly expressed in the brain, and more than 3600 brain-specific lncRNAs have been identified, and they play a role in neural development, proliferation, and differentiation as epigenetic regulators [188]. lncRNAs are involved in the onset and development of schizophrenia, MDD, bipolar depression, and suicide [189]. lncRNAs affect depression by interacting with the epigenetic modification of DNA and chromatin, gene SNPs, and compete with endogenous RNA networks related to NTF expression and synaptic function [190].
Epidemiological and preclinical studies have presented that early-life child maltreatment, parent neglect, undernutrition, or sexual abuse increase the risk for depression and other stress-related disorders by two- to four-fold [191,192]. An unfavorable maternal environment induces long-term epigenomic changes in gene expression related to 5-HT that persist into adulthood [193]. In adult males with high childhood-limited aggression, early-life stress promoted methylation levels of the SLC6A4 promoter in T cells and monocyte, and associated with lower 5-HT synthesis in the lateral orbitofrontal cortex measured in vivo using PET imaging [194]. Increased DNA methylation in the promoter region of 5-HTR1A gene was observed in schizophrenia and bipolar disorder [195]. A study on the 5-HTT methylation status in 10-year-old monozygotic twins presented that 5-HTT levels were significantly higher in the twin who experienced bullying and stress than the one who did not undergo such experiences [196]. MAO-A expression and activity are negatively correlated with the methylation status of the MAO-A promotor at CpG islands (CGIs) including 14 CpG sites [197].
Sulforaphane, flavonoids, resveratrol, and curcumin can reverse abnormal gene expression by epigenomic mechanisms, and exert therapeutic potential for neuropsychiatric disorders, aging, cancer, and chronic inflammation [198,199,200]. The impact of diets on epigenetic regulation is mainly the supplementation of methyl groups from methionine, vitamin B12, folate, betaine (trimethylglycine), and polyphenols (genistein) [201]. Curcumin, genistein, apigenin, luteolin, catechins, and EGCG activate NF-κB expression, inhibit DNMT and HDAC, and remodulate chromatin conformation and reverse abnormal gene expression. A hydroxyl group at the C-7 position in ring B of fisetin (3,7,3′,4′-tetrahydroxyflavone), silibinin, and daidzein are required for the activation of DNMT and SIRT1, whereas a hydroxyl group at the same position in luteolin and EGCG inhibits SIRT1. Genistein, myricetin, and quercetin with hydroxyl groups at positions C-5 and C-7, inhibit or activate HDACs and DNMTs depending on experimental conditions [202]. Quercetin, icariin, silibinin, daidzein, formonetin, and biochain A increased SIRT1 and PGC-1α expression, increased AMPK phosphorylation, induced histone deacetylation, and exerted neuroprotection in cellular models of neurodegenerative diseases. Polyphenolic HDAC activators, such as quercetin, halted or delayed the propagation of aging and significantly prolonged the life span of mice through the inhibition of SIRT1 by quercetin-3-O-gluconide, a quercetin metabolite [203]. Curcumin increased miR-128 and miR-9 and downregulated phosphorylated tau in rat cortical neurons [204]. Apigen increased miR-15a expression, decreased Rho-associated protein kinase-1 (ROCK-1), and exerted neuroprotection in the rat hippocampus [205]. Genistein and E2 enhanced miR-132, BDNF, and IGF-1 in the hippocampus of ovariectomized rats and improved spatial memory [206]. Resveratrol downregulated lncRNAs (SNHG1, MALAT1), and upregulated the miRNAs (miR-361 -3p, miR-129) and TH cells in cell models of AD and PD [207,208]. Berberine downregulated BACE1-AS and LINC00943, increased miR-129 and miR-142n, and showed neuroprotection and anti-inflammation and anti-apoptosis activity [209]. Ginsenoside Rf modulated the expression of lncRNAs (Mett127, Cyp23, MSN, Ptx3, et al.) and inhibited tau aggregation and toxicity [210].
Epigenomic phytochemicals have been clinically tried as attractive therapeutic agents for cancer, aging, cognitive decline, obesity, and chronic inflammation [211,212]. In healthy premenopausal women, isoflavones (40 or 140 mg, daily, through one menstrual cycle) increased the methylation of breast cancer-related genes RARb2 and CCND2 in correlation with serum genistein levels [213]. The administration of genistein (54 mg daily for 2 years) in postmenopausal women decreased the depression score, which was assessed with the Zung Self-rating Depression Scale [214]. Polyphenols, saponins and terpenoids, and alkaloids have been proposed as antidepressant candidates without severe side-effects by the upregulation of 5-HT, NE and DA, BDNF induction, MAO inhibition, and the suppression of HPA axis overactivity [215]. The efficacy of phytochemicals as antidepressants has been presented by preclinical investigations and some clinical investigations [216]. Curcumin administration reduced depressive symptoms in patients with MDD [217].

8. Clinical Applications of Phytochemical Treatment in Neuropsychiatric Disorders

The malfunction of the 5-HT system, including TPH, 5-HT1, 5-HT7, and SERT, is involved in mood disorders (depression, anxiety), cognition decline, and neurodevelopmental disorders. There is a critical window in a distinct stage of brain development, during which 5-HT levels permanently regulate neuronal architecture [218]. Can phytochemicals prevent or ameliorate the 5-HT-related impairment of brain development during early stage? Preclinical evidence suggests that dietary and bioactive phytochemicals can influence monoamine biosynthesis, metabolism, and function, with certain compounds, particularly phytoestrogens, also exhibiting estrogen- and BDNF-like activities. These properties may contribute to their neuroprotective effects via epigenetic modifications, highlighting their potential role in modulating neurodevelopment and neurotransmitter function. Many neuroprotective phytochemicals, including alkaloids, polyphenols, and terpenoids, have demonstrated promising in vivo and in vitro effects but remain largely unexplored in humans. Recent meta-analyses have highlighted several important compounds, including huperzine A, caffeine, curcumin, resveratrol, quercetin, ginkgolide, ferulic acid, rosmarinic acid, as potential therapeutic agents across conditions such as depression, anxiety, schizophrenia, AD, and dementia [25,219,220,221] These compounds have progressed to clinical trials [221,222].
While these findings support their potential as therapeutic agents for monoamine-related diseases, direct clinical evaluations remain limited due to multiple challenges in clinical application. Neuropsychiatric disorders involve complex, multifactorial systems, making it difficult to establish clear therapeutic targets. Additionally, the diverse and heterogeneous nature of these disorders contributes to inconsistent clinical outcomes, sometimes yielding contradictory results that require more reliable and reproducible evidence [223,224]. Accessibility issues, ambiguous effectiveness, and potential side effects further complicate their translation into clinical use [225]. Moreover, phytochemicals are frequently administered as adjunctive or complementary therapies, often in the form of plant extracts or multi-component formulations, making it challenging to isolate and assess the effects of individual compounds [226,227]. Alongside these factors, poor bioavailability, limited blood-brain barrier permeability, and variability in assessment methods further hinder their clinical validation [228].
One problem is how to effectively evaluate the effects of phytochemicals on monoamine systems in vivo. Monoamine neurotransmitter concentrations can be measured in cerebrospinal fluid (CSF) to monitor their synthesis, metabolism, and the activity of the related enzymes. The levels and activity of monoamine transmitters can be indirectly visualized using PET or single-photon emission computed tomography (SPECT) [229]. For example, the 5-HT levels in children with autism could be determined by measuring SERT expression using SPECT with [131I]-N-(2-fluoroethyl)-2b-carbomethoxy-3b-(4-iodophenyl)-nortropane [230]. In humans, MAO level and activity are measured using PET imaging with [11C]- or [18F]-radiotracer aliphatic amines and MAO inhibitors, commonly harmine [231,232]. These advanced imaging techniques offer a powerful, non-invasive approach to monitoring monoamine neurotransmission; however, optimizing their cost, accessibility, and technical feasibility is essential for broader clinical application.
Another key limitation in the clinical application of phytochemicals for neuropsychiatric diseases is their poor bioavailability and stability due to extensive metabolism. In the digestive system, phytochemicals undergo hydrolysis, microbial degradation, hepatic metabolism, and rapid excretion, significantly reducing their systemic availability. In humans, only a small part of ingested polyphenols (0.3–43%) are found in urine, indicating their poor bioavailability [233]. To overcome this, various nanotechnology-based delivery systems, such as solid lipid nanoparticles (SLNs), chitosan nanoparticles, and polymeric nanocarriers, have been developed to enhance absorption, improve stability, and promote CNS targeting for compounds like genistein and curcumin [234,235]. Additionally, alternative administration routes, including the intranasal delivery of genistein-loaded chitosan nanoparticles and transdermal administration of resveratrol-loaded SLNs (RSV-SLNs) via microneedle patches, have been explored to bypass first-pass metabolism and enhance systemic retention [235,236,237]. Prodrug approaches, such as the conversion of curcumin to a curcumin diethyl γ-aminobutyrate (CUR-2GE) prodrug has been shown to enhance its bioavailability and anti-neuroinflammatory properties [238]. Curcumin–piperine-loaded self-nanoemulsifying drug delivery systems (SNEDDS) have presented a safe and effective oral delivery method for improving curcumin bioavailability and brain targeting in AD models [239].
Once phytochemicals achieve sufficient systemic circulation, the next challenge is crossing the BBB to exert neuropsychiatric effects. The restricted permeability of the BBB limits the therapeutic efficacy of many phytochemicals, as only compounds with optimal lipophilicity, molecular size, and transport mechanisms can effectively reach neuronal targets [240] The brain entry of polyphenols depends on their stereochemistry, lipophilicity, and interactions with efflux transporters, such as P-glycoprotein (PGP) [241]. Many flavonoids, flavones, flavonols, isoflavones, phenolic acids, and stilbenes are ligands for brain efflux transporters and penetrate the brain to varying degrees [242]. Lipophilic modifications, such as O-methylation and esterification, enhance BBB penetration [243]. Lipid-based and mucoadhesive systems, including glucose-modified liposomes targeting the glucose transporter 1 (GLUT1), have demonstrated increased BBB permeability and neuroprotection of quercetin in oxidative stress models [244]. Curcumin-loaded exosomes have been investigated as a strategy to facilitate BBB crossing, enhancing its neuroprotective effects [245]. Additionally, borneol, a natural monoterpene, has been investigated for its ability to transiently increase BBB permeability, thereby improving the brain bioavailability of co-administered neuroprotective agents. It modulates efflux transporters, alters tight junction proteins, and enhances vasodilatory neurotransmitters, facilitating a greater CNS uptake of therapeutic compounds [246].
While phytochemicals hold therapeutic potential in neuropsychiatric diseases, their clinical application remains limited by many challenges. Advances in formulation strategies, biotechnologies, and permeability enhancers like borneol provide promising solutions to these limitations, but research has largely focused on a few well-studied compounds like curcumin, resveratrol, and quercetin. Expanding research efforts and clinical validation to a wider range of phytochemicals is crucial to fully realize their therapeutic potential in monoamine-related psychiatric diseases.

9. Conclusions

This review explored the role of phytochemicals in modulating 5-HT, DA, NE biosynthesis, metabolism, and function, highlighting their potential in treating neuropsychiatric disorders. While challenges such as limited bioavailability and BBB permeability remain, advances in nanotechnology and targeted delivery systems offer promising solutions. Notably, phytochemicals and their derivatives act as potent regulators of MAO-A, which, due to its flexibility to environmental changes, plays a more dynamic role in maintaining monoamine neurotransmission than THP and TH [77]. Beyond therapy, the intervention of phytochemicals in the 5-HT-dependent regulation of early brain development may be a novel strategy for prevention and treatment psychiatric disorders, including aggression, depressive disorders, and anxiety.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms26072916/s1.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

AP-1activating protein-1
BH4(6R)-L-tetrahydrobiopterin
CREBCRE binding protein
ΔΨmmitochondrial membrane potential
E217β-estradiol
ERestrogen receptor
EREestrogen response element
GTPCHguanidine triphosphate cyclohydrolase
mPTPmitochondrial permeability transition pore
NTFneurotrophic factor
Nrf2nuclear factor erythroid 3-related factor
PGC-1αperoxisome proliferator-activated receptor γ coactivator α
SERMER type-specific ER modulators
SERT5-HT transporter
Sp1simian virus 40 promoter factor 1
SRYsex-determining region Y

References

  1. Whitaker-Azmitia, P.M. Serotonin and brain development: Role in human developmental diseases. Brain Res. Bull. 2001, 56, 479–485. [Google Scholar] [CrossRef] [PubMed]
  2. Lauder, J.M. Serotonin as a differentiation signal. Dev. Neurosci. 2016, 38, 239–240. [Google Scholar] [CrossRef] [PubMed]
  3. Lesch, K.P.; Waider, J. Serotonin in the modulation of neural plasticity and networks: Implications for neurodevelopmental disorders. Neuron 2012, 76, 175–191. [Google Scholar] [CrossRef]
  4. 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] [PubMed]
  5. Burgoyne, R.D.; Morgan, A. Secretory granule exocytosis. Physiol. Rev. 2003, 83, 581–632. [Google Scholar] [CrossRef]
  6. Aurora, M.; Per, M.K.; Jan, H. A Structural Approach into Human Tryptophan Hydroxylase and its Implications for the Regulation of Serotonin Biosynthesis. Curr. Med. Chem. 2001, 8, 1077–1091. [Google Scholar] [CrossRef]
  7. Fernstrom, J.D.; Fernstrom, M.H. Tyrosine, phenylalanine, and catecholamine synthesis and function in the brain. J. Nutr. 2007, 137 (Suppl. S1), 1539S–1547S. [Google Scholar] [CrossRef]
  8. Dunkley, P.R.; Bobrovskaya, L.; Graham, M.E.; von Nagy-Felsobuki, E.I.; Dickson, P.W. Tyrosine hydroxylase phosphorylation: Regulation and consequence. J. Neurochem. 2004, 91, 1025–1043. [Google Scholar] [CrossRef]
  9. Kumer, S.C.; Vrana, K.E. Intricate regulation of tyrosine hydroxylase activity and gene expression. J. Neurochem. 1996, 67, 443–462. [Google Scholar] [CrossRef]
  10. Chen, G.L.; Miller, G.M. Advances in tryptophan hydroxylase-2 gene expression regulation: New insights into serotonin-stress interaction and clinical implication. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2012, 159B, 152–171. [Google Scholar] [CrossRef]
  11. Bendis, P.C.; Zimmerman, S.; Onisiforou, A.; Zanos, P.; Georgiou, P. The impact of estradiol on serotonin, glutamate, and dopamine systems. Front. Neurosci. 2024, 18, 1348551. [Google Scholar] [CrossRef]
  12. Ferrnstrom, J.D. Effects on the diet on brain neurotransmitters. Metabolism 1977, 26, 207–223. [Google Scholar] [CrossRef] [PubMed]
  13. Ki, M.R.; Youn, S.; Kim, D.H.; Pack, S.P. Natural compounds for preventing age-related diseases and cancers. Int. J. Mol. Sci. 2024, 25, 7530. [Google Scholar] [CrossRef] [PubMed]
  14. Singla, R.K.; Dubey, A.K.; Garg, A.; Sharma, R.K.; Fiorino, M.; Ameen, S.M.; Haddad, M.A.; Al-Hiary, M. Natural Polyphenols: Chemical Classification, Definition of Classes, Subcategories, and Structures. J. AOAC Int. 2019, 102, 1397–1400. [Google Scholar] [CrossRef]
  15. Rebas, E.; Rzajew, J.; Radzik, T.; Zylinska, L. Neuroprotective polyphenols: A modulatory action on neurotransmission pathways. Curr. Neuropharmacol. 2020, 18, 431–445. [Google Scholar] [CrossRef]
  16. Naoi, M.; Maruyama, W.; Shamoto-Nagai, M. Disease-modifying treatment of Parkinson’s disease by phytochemicals: Targeting multiple pathogenic factors. J. Neural Transm. 2022, 129, 737–753. [Google Scholar] [CrossRef] [PubMed]
  17. Noori, T.; Sureda, A.; Sobarzo-Sanchez, E.; Shirooie, S. The role of natural products in treatment of depressive disorder. Curr. Neuropharmacol. 2022, 20, 929–949. [Google Scholar] [CrossRef]
  18. Gorzkiewicz, J.; Bartosz, G.; Sadowska-Bartosz, I. The potential effects of phytoestrogens: The role in neuroprotection. Molecules 2021, 26, 2954. [Google Scholar] [CrossRef]
  19. Yanagihara, N.; Zhang, H.; Toyohira, Y.; Takahashi, R.; Ueno, S.; Tsutsui, M.; Takahashi, K. New insight into the pharmacological potential of plant flavonoids in the catecholamine system. J. Pharnacol. Sci. 2014, 124, 123–128. [Google Scholar] [CrossRef]
  20. Shah, R.; Courtiol, E.; Castellanos, F.X.; Teixeira, C.M. Abnormal serotonin levels during perinatal development lead to behavioral deficits in adulthood. Front. Behav. Neurosci. 2018, 12, 114. [Google Scholar] [CrossRef]
  21. Pandareesh, M.D.; Mythri, R.B.; Srinivas Bharath, M.M. Bioavailability of dietary polyphenols: Factors contributing to their clinical application in CNS diseases. Neurochem. Int. 2015, 89, 198–208. [Google Scholar] [CrossRef] [PubMed]
  22. Naoi, M.; Inaba-Hasegawa, K.; Shamoto-Nagai, M.; Maruyama, W. Neurotrophic function of phytochemicals for neuroprotection in aging and neurodegenerative disorders: Modulation of intracellular signaling and gene expression. J. Neural Transm. 2017, 124, 1515–1527. [Google Scholar] [CrossRef]
  23. Naoi, M.; Wu, Y.; Shamoto-Nagai, M.; Maruyama, W. Mitochondrial in neuroprotection by phytochemicals: Bioactive polyphenols modulate mitochondrial apoptosis system, function and structure. Int. J. Mol. Sci. 2019, 20, 2451. [Google Scholar] [CrossRef]
  24. Wu, Y.; Shamoto-Nagai, M.; Maruyama, W.; Osawa, T.; Naoi, M. Phytochemicals prevent mitochondrial membrane permeabilization and protect SH-SY5Y cells against apoptosis induced by PK11195, a ligand for outer membrane translocator protein. J. Neural Transm. 2017, 124, 89–98. [Google Scholar] [CrossRef] [PubMed]
  25. Asgharian, P.; Quispe, C.; Herrea-Bravo, J.; Sabernavaei, M.; Hosseini, K.; Forouhandeh, H.; Ebrahimi, T.; Sharafi-Badr, P.; Tarhriz, V.; Soofiyani, S.R.; et al. Pharmacological effects and therapeutic potential of natural compounds in neuropsychiatric disorders: An update. Front. Pharmacol. 2022, 13, 926607. [Google Scholar] [CrossRef]
  26. Fitzpatrick, P.F. Mechanism of aromatic amino acid hydroxylation. Biochemistry 2003, 42, 14083–14091. [Google Scholar] [CrossRef]
  27. Walther, D.J.; Baker, M. A unique central tryptophan hydroxylase isoform. Biochem. Pharmacol. 2003, 66, 1673–1680. [Google Scholar] [CrossRef]
  28. Murphy, K.L.; Zhang, X.; Gainetdinov, R.R.; Beaulieu, J.M.; Caron, M.G. A regulatory domain in the N terminus of tryptophan hydroxylase 2 controls enzyme expression. J. Biol. Chem. 2008, 283, 13216–13224. [Google Scholar] [CrossRef]
  29. Zill, P.; Baghai, T.C.; Zwanzger, P.; Schüle, C.; Eser, D.; Ruppercht, R.; Möller, H.J.; Bondy, B.; Ackenheil, M. SNP and haplotype analysis of a novel tryptophan isoform (TPH2) gene provide evidence for association with major depression. Mol. Psychiatry 2004, 9, 1030–1036. [Google Scholar] [CrossRef]
  30. Cichon, S.; Winge, I.; Mattheisen, M.; Georgi, A.; Karpushova, A.; Freudenberg, J.; Freudenberg-Hua, Y.; Babadjanova, G.; Van Den Bogaert, A.; Abramova, L.I.; et al. Brain-specific tryptophan hydroxylase 2 (TPH2): A functional Pro206Ser substitution and variation in the 5′-region are associated with bipolar affective disorder. Hum. Mol. Genet. 2008, 17, 87–97. [Google Scholar] [CrossRef]
  31. Zhang, X.; Beaulieu, J.M.; Gainetdinov, R.R.; Caron, M.G. Functional polymorphisms of the brain serotonin synthesizing enzyme tryptophan hydroxylase-2. Cell. Mol. Life Sci. 2006, 63, 6–11. [Google Scholar] [CrossRef] [PubMed]
  32. Bademci, G.; Vance, J.M.; Wang, L. Tyrosine hydroxylase gene: Another piece of the genetic puzzle of Parkinson’s disease. CNS Neurol. Disord. Drug Targets 2012, 11, 469–481. [Google Scholar] [CrossRef] [PubMed]
  33. Daubner, S.C.; Le, T.; Wang, S. Tyrosine hydroxylase and regulation of dopamine synthesis. Arch. Biochem. Biophys. 2011, 508, 1–12. [Google Scholar] [CrossRef] [PubMed]
  34. Ramsey, A.J.; Fitzpatrick, P.F. Effects of phosphorylation of serine 40 of tyrosine hydroxylase on binding of catecholamines: Evidence of a novel regulatory mechanism. Biochemistry 1998, 37, 8980–8986. [Google Scholar] [CrossRef]
  35. Nagatsu, T.; Nakshima, A.; Ichinose, H.; Kobayashi, K. Juman tyrosine hydroxylase in Parkinson’s disease and in related disorders. J. Neural Transm. 2019, 126, 397–409. [Google Scholar] [CrossRef]
  36. Willemsen, M.A.; Verbeek, M.M.; Kamsteeg, E.J.; de Rijk-van Andel, J.F.; Aeby, A.; Blau, N.; Burlina, A.; Donati, M.A.; Geurtz, B.; Grattan-Smith, P.J.; et al. Tyrosine hydroxylase deficiency: A treatable disorder of brain catecholamine biosynthesis. Brain 2010, 1333 Pt 6, 1810–1822. [Google Scholar] [CrossRef]
  37. Nagatomo-Combs, K.; Piech, K.M.; Best, J.A.; Sun, B.; Tank, A.W. Tyrosine hydroxylase gene promoter activity is regulated by both cyclic AMP-responsive element and AP1 sites following calcium influx. J. Biol. Chem. 1997, 272, 6051–6058. [Google Scholar] [CrossRef]
  38. Sabban, E.L. Control of tyrosine hydroxylase gene expression in chromatin and PC12 cells. Semin. Cell Dev. Bilo. 1997, 8, 101–111. [Google Scholar] [CrossRef]
  39. Montioli, R.; Voltattorni, C.B. Aromatic amino acid decarboxylase deficiency: The added value of biochemistry. Int. J. Mol. Sci. 2021, 22, 3146. [Google Scholar] [CrossRef]
  40. Sumi-Ichinose, C.; Ichinose, H.; Takahashi, E.; Hori, T.; Nagatsu, T. Molecular cloning of genomic DNA and chromosomal assignment of the gene for human aromatic L-amino acid decarboxylase, the enzyme for catecholamine and serotonin synthesis. Biochemistry 1992, 31, 2229–2238. [Google Scholar] [CrossRef]
  41. Rizzi, S.; Spagnoli, C.; Frattini, D.; Pisani, F.; Fusco, C. Clinical features in aromatic L-amino acid decarboxylase (AADC) deficiency: A systematic review. Behav. Neurol. 2022, 2022, 2210555. [Google Scholar] [CrossRef] [PubMed]
  42. Shih, J.C. Monoamine oxidase isoenzymes: Genes, functions and targets for behavior and cancer therapy. J. Neural Transm. 2018, 125, 1553–1566. [Google Scholar] [CrossRef]
  43. Naoi, M.; Riederer, P.; Maruyama, W. Modulation of monoamine oxidase (MAO) expression in neuropsychiatric disorders: Genetic and environmental factors involved in type A MAO expression. J. Neural Transm. 2016, 123, 91–106. [Google Scholar] [CrossRef]
  44. Zhu, Q.; Shih, J.C. An extensive repeat structure down-regulates human monoamine oxidase A promoter activity independent of an initiator-like sequence. J. Neurochem. 1997, 69, 1368–1373. [Google Scholar] [CrossRef] [PubMed]
  45. Contini, V.; Marques, F.Z.C.; Garcia, C.E.D.; Hutz, M.H.; Bau, C.H.D. MAOA-uVNTR polymorphism in a Brazilian sample: Further support for the association with impulsive behaviors and alcohol dependence. Am. J. Med. Gent. B. Neuropsychiatr. Genet. 2006, 141B, 305–308. [Google Scholar] [CrossRef]
  46. Fan, M.; Liu, B.; Jiang, T.; Jiang, X.; Zhao, H.; Zhang, J. Meta-analysis of the association between the monoamine oxidase-A gene and mood disorders. Psychiatr. Genet. 2010, 20, 1–7. [Google Scholar] [CrossRef] [PubMed]
  47. Brunner, H.G.; Nelen, M.; Breakfield, X.O.; Ropers, H.H.; van Oost, B.A. Abnormal behavior associated with a point mutation in the structural gene for monoamine oxidase A. Science 1991, 262, 578–580. [Google Scholar] [CrossRef]
  48. Lenders, J.W.; Eisenhofer, G.; Abeling, N.G.; Berger, W.; Murphy, D.L.; Konings, C.H.; Wagemakers, L.M.; Kopin, I.J.; Karoum, F.; van Gennip, A.H.; et al. Specific genetic deficiencies of the A and B isoenzymes of monoamine oxidase are characterized by distinct neurochemical and clinical phenotypes. J. Clin. Investig. 1996, 97, 1010–1019. [Google Scholar] [CrossRef]
  49. Matsumoto, M.; Shannon Weickert, C.; Akil, M.; Lipska, B.K.; Hyde, T.M.; Herman, M.M.; Kleinman, J.E.; Weinberger, D.R. Catechol O-methyltransferase mRNA expression in human and rat brain: Evidence for a role in cortical neuronal function. Neuroscience 2003, 116, 127–137. [Google Scholar] [CrossRef]
  50. Harrison, P.J.; Tunbridge, E.M. Catechol-O-methyltransferase: A gene contributing to sex differences in brain function, and to sexual dimorphism in the preposition to psychiatric disorders. Neuropsychopharmacology 2008, 33, 3037–3045. [Google Scholar] [CrossRef]
  51. Bastos, P.; Gomes, T.; Ribeiro, L. Catechol-O-methyltransferase: An update on its role in cancer, neurological and cardiovascular diseases. Rev. Physiol. Biochem. Pharmacol. 2017, 173, 1–39. [Google Scholar] [CrossRef] [PubMed]
  52. Clelland, J.D.; Hesson, H.; Ramiah, K.; Anderson, J.; Thengampallil, A.; Girgis, R.R.; Chelland, C.L. The relationship between COMT, proline, and negative symptoms in clinical high risk and recent psychosis onset. Transl. Psychiatry 2024, 14, 409. [Google Scholar] [CrossRef] [PubMed]
  53. Babazadeh, A.; Vahed, F.M.; Liu, Q.; Siddiqui, S.A.; Khazazmi, M.S.; Jafari, S.M. Natural bioactive molecules as neuromedicines for the treatment/prevention of neurodegenerative diseases. ACS Omega 2023, 8, 3667–3683. [Google Scholar] [CrossRef]
  54. Vauzour, D.; Rodriguez-Mateos, A.; Corona, G.; Oruna-Concha, M.J.; Spencer, J.P.E. Polyphenols and human health: Prevention of disease and mechanisms of action. Nutrients 2010, 2, 1106–1131. [Google Scholar] [CrossRef] [PubMed]
  55. Cheynier, V. Polyphenols in foods are more complex than often thought. Am. J. Clin. Nutr. 2005, 81, 223S–229S. [Google Scholar] [CrossRef]
  56. Perron, N.R.; Brumahim, J.L. A review of the antioxidant mechanisms of polyphenol compounds related to iron binding. Cell. Biochem. Biophys. 2009, 53, 75–100. [Google Scholar] [CrossRef]
  57. Fraga, C.G.; Galleano, M.; Verstraeten, S.V.; Oteiza, P.I. Basic biochemical mechanisms behind the health benefits of polyphenols. Mol. Aspects. Med. 2010, 31, 435–445. [Google Scholar] [CrossRef]
  58. Leonardo, C.C.; Dore, S. Dietary flavonoids are neuroprotective through Nrf2-coordinated induction of endogenous cytoprotective proteins. Nutr. Neurosci. 2011, 14, 226–236. [Google Scholar] [CrossRef]
  59. Habtemariam, S. The Nrf2/HO-1 axis as targets for flavanones: Neuroprotection by pinocembrin, naringenin, and eriodictyol. Oxid. Med. Cell Longev. 2019, 13, 4724920. [Google Scholar] [CrossRef]
  60. Williams, R.J.; Spencer, J.P.E.; Rice-Evans, C. Flavonoids: Antioxidants or signaling molecules? Free Rad. Biol. Med. 2004, 36, 838–849. [Google Scholar] [CrossRef]
  61. Spencer, J.P.E. The interactions of flavonoids within neuronal signaling pathways. Genes Nutr. 2007, 2, 257–273. [Google Scholar] [CrossRef] [PubMed]
  62. Moosavi, F.; Hosseini, R.; Saso, L.; Firuz, O. Modulation of neurotrophic signaling pathways by polyphenols. Drug Des. Devel. Ther. 2015, 10, 23–42. [Google Scholar] [CrossRef]
  63. Venkatesan, R.; Ji, E.; Kim, S.Y. Phytochemicals that regulate neurodegenerative disease by targeting neurotrophins: A comprehensive review. Biomed. Res. Int. 2015, 2015, 814068. [Google Scholar] [CrossRef] [PubMed]
  64. Neshatdoust, S.; Saunders, C.; Castle, S.M.; Vauzour, D.; Williams, C.; Butler, L.; Lovegrove, J.A.; Spencer, J.P. High-flavonoid intake induced cognitive improvement linked to changes in serum brain-derived neurotrophic factor: Two randomized, controlled trials. Nutr. Healthy Aging 2016, 4, 81–93. [Google Scholar] [CrossRef]
  65. Fanaei, H.; Khayat, S.; Kasaeian, A.; Jvadimehr, M. Effect of curcumin on serum brain-derived neurotrophic factor levels in women with premenstrual syndrome: A randomized, double-blind, placebo-controlled trial. Neuropeptides 2016, 56, 25–31. [Google Scholar] [CrossRef] [PubMed]
  66. Kessas, K.; Chouari, Z.; Ghzaiel, I.; Zarrouk, A.; Ksila, M.; Ghrairi, T.; El Midaoui, A.; Lizard, G.; Kharoubi, O. Role of bioactive compounds in the regulation of mitochondrial dysfunctions in the brain and age-related neurodegenerative diseases. Cells 2022, 11, 257. [Google Scholar] [CrossRef]
  67. Lagouge, M.; Argann, C.; Gerhart-Hines, Z.; Meziane, H.; Lerin, C.; Daussin, F.; Messadeq, N.; Milne, J.; Lambert, P.; Elliott, P.; et al. Resveratrol improves mitochondrial function and protects against metabolic disease by activating SIRT1 and PGC-1α. Cell 2006, 127, 1109–1122. [Google Scholar] [CrossRef]
  68. Nagley, P.; Higgins, G.C.; Atkin, J.D.; Beart, P.M. Multifaceted death orchestrated by mitochondria in neurons. Biochim. Biophys. Acta 2010, 1802, 167–185. [Google Scholar] [CrossRef]
  69. Wu, Y.; Kazumura, K.; Maruyama, W.; Osawa, T.; Naoi, M. Rasagiline and selegiline suppress calcium efflux from mitochondria by PK11195-induced opening of mitochondrial permeability transition pore: A novel anti-apoptotic function for neuroprotection. J. Neural Transm. 2015, 122, 1399–1407. [Google Scholar] [CrossRef]
  70. Wu, Y.; Shamoto-Nagai, M.; Maruyama, W.; Osawa, T.; Naoi, M. Rasagiline prevents cyclosporine A-sensitive superoxide flashes induced by PK11195, the initial signal of mitochondrial membrane permeabilization and apoptosis. J. Neural Transm. 2016, 123, 491–494. [Google Scholar] [CrossRef]
  71. Krestinina, O.; Baburina, Y.; Krestinin, R.; Odinokova, I.; Fedeeva, I.; Sotnikowa, L. Astaxanthin prevents mitochondrial impairment induced by isoproterenol in isolated rat heart mitochondria. Antioxidants 2020, 9, 262. [Google Scholar] [CrossRef] [PubMed]
  72. Xi, J.; Wang, H.; Müller, R.M.; Nirfleet, E.A.; Xu, Z. Mechanism for resveratrol-induced cardioprotection against reperfusion injury involves glycogen synthase kinas 3b and mitochondrial permeability transition pore. Eur. J. Pharmacol. 2009, 604, 111–116. [Google Scholar] [CrossRef] [PubMed]
  73. Liao, Z.; Liu, D.; Tang, L.; Yin, D.; Yin, S.; Yao, J.; He, M. Long-term oral resveratrol intake provides nutritional preconditioning against myocardial ischemia/reperfusion injury: Involvement of VDAC1 downregulation. Mol. Nutr. Food Res. 2015, 59, 454–464. [Google Scholar] [CrossRef]
  74. Tian, M.; Xie, Y.; Meng, Y.; Ma, W.; Tong, Z.; Yang, X.; Lai, S.; Zhou, Y.; He, M.; Liao, Z. Resveratrol protects cardiomyocytes against anoxia/reoxygenation via dephosphorylation of VDAC1 by Akt-GSK3 b pathway. Eur. J. Pharmcol. 2019, 843, 80–87. [Google Scholar] [CrossRef]
  75. Tewari, D.; Ahmed, T.; Chrirasani, V.R.; Singh, P.K.; Maji, S.M.; Senapati, S.; Bera, A.K. Modulation of the mitochondrial voltage dependent anion channel (VDAC) by curcumin. Biochim. Biophys. Acta. 2015, 1848, 151–158. [Google Scholar] [CrossRef]
  76. De Marchi, U.; Bissutto, L.; Garbisa, S.; Toninello, A.; Zoratti, M. Quercetin can act an inhibitor or an inducer of the mitochondrial permeability transition pore: A demonstration of the ambivalent redox character of polyphenols. Biochim. Biophys. Acta 2009, 1787, 1425–1432. [Google Scholar] [CrossRef]
  77. Naoi, M.; Maruyama, W.; Shamoto-Nagai, M.; Riederer, P. Type A monoamine oxidase; its unique role in mood, behavior and neurodegeneration. J. Neural Transm. 2024, in press. [CrossRef] [PubMed]
  78. Sabban, E.L.; Maharjan, S.; Nistramo, R.; Serova, L.I. Divergent effects of estradiol on gene expression of catecholamine biosynthetic enzymes. Physiol. Behav. 2010, 99, 163–168. [Google Scholar] [CrossRef]
  79. Hernandez-Hernandez, O.T.; Martinez-Mota, L.; Herrera-Perez, J.; Jimenez-Rubio, G. Role of estradiol in the expression of genes involved in serotonin neurotransmission: Implications for female depression. Curr. Neuropharmacol. 2019, 17, 459–471. [Google Scholar] [CrossRef]
  80. De Vries, G.J. Minireview: Sex differences in adult and developing brains: Compensation, compensation, compensation. Endocrinology 2004, 145, 1063–1068. [Google Scholar] [CrossRef]
  81. Sumien, N.; Cunnungham, J.T.; Davis, D.L.; Engelland, R.; Fadeyibi, O.; Farmer, G.E.; Mabry, S.; Mensah-Kane, P.; Trinh, O.T.P.; Vann, P.H.; et al. Neurodegenerative disease: Role for sex, hormones, and oxidative stress. Endocrinology 2021, 162, bdab185. [Google Scholar] [CrossRef] [PubMed]
  82. Voskuhl, R.; Itoh, Y. The X factor in neurodegeneration. J. Exp. Med. 2022, 219, e20211488. [Google Scholar] [CrossRef] [PubMed]
  83. Chen, P.; Li, B.; Ou-Yang, L. Role of estrogen receptors in health and disease. Front. Endocrinol. 2022, 13, 839005. [Google Scholar] [CrossRef]
  84. Hall, J.M.; Couse, J.F.; Korach, K.S. The multi faced mechanisms of estradiol and estrogen receptor signaling. J. Biol. Chem. 2001, 276, 36869–36872. [Google Scholar] [CrossRef]
  85. Nilsson, S.; Mäkelä, S.; Treuter, E.; Tujague, M.; Thomsen, J.; Andersson, G.; Enmark, E.; Pettersson, K.; Warner, M.; Gustafsson, J.A. Mechanisms of estrogen action. Physiol. Rev. 2001, 81, 1535–1565. [Google Scholar] [CrossRef]
  86. Wissink, S.; van der Burg, B.; Katzenellenbogen, B.S.; van der Saag, P.T. Synergistic activation of the serotonin-1A receptor by nuclear factor-kB and estrogen. Mol. Endocrinol. 2001, 15, 543–552. [Google Scholar] [CrossRef]
  87. Safe, S.; Kim, K. Non-classical genomic estrogen receptor (ER)/specificity protein and ER/activating protein-1 signaling pathway. J. Mol. Endocrinol. 2008, 41, 263–275. [Google Scholar] [CrossRef]
  88. Gundlah, C.; Lu, N.Z.; Mirkes, S.J.; Bethea, S.L. Estrogen receptor b (ERb) mRNA and protein in serotonin neurons of macaques. Brain Res. Mol. Brain Res. 2001, 91, 14–22. [Google Scholar] [CrossRef]
  89. Hudon Thibeault, A.A.T.; Sanderson, J.T.; Vaillamcourt, C. Serotonin-estrogen interactions: What can we learn from pregnancy? Biochimie 2019, 161, 88–108. [Google Scholar] [CrossRef]
  90. Hiroi, R.; Handa, R.J. Estrogen receptor-b regulates human tryptophan hydroxylase-2 through an estrogen response element in the 5′ untranslated region. J. Neurochem. 2013, 127, 487–495. [Google Scholar] [CrossRef]
  91. Sanchez, R.L.; Reddy, A.P.; Centeno, M.L.; Henderson, J.A.; Betha, C.L. A second tryptophan hydroxylase isoform, TPH-2 mRNA, is increased by ovarian steroids in the raphe regions of macaques. Brain Res. Mol. Brain Res. 2005, 135, 194–203. [Google Scholar] [CrossRef]
  92. Hiroi, R.; McDevitt, R.A.; Neumaler, J.F. Estrogen selectively increases tryptophan hydroxylase-2 mRNA expression in distinct subregions of rat midbrain raphe nucleus: Association between gene expression and anxiety behavior in the open field. Biol. Psychiatry 2006, 60, 288–295. [Google Scholar] [CrossRef] [PubMed]
  93. Imwalle, D.B.; Gustafssom, J.A.; Rissman, E.F. Lack of functional estrogen receptor β influences anxiety behavior and serotonin content in female mice. Physiol. Behav. 2005, 84, 157–163. [Google Scholar] [CrossRef]
  94. Nomura, M.; Akama, K.T.; Alves, S.E.; Korach, K.S.; Gustafsson, J.A.; Pfaff, D.W.; Ogawa, S. Differential distribution of estrogen receptor (ER)-a and ER-b in the midbrain raphe nuclei and periaqueductal gray in male mouse: Predominant role of ER-b in the midbrain serotonergic system. Neuroscience 2005, 130, 445–456. [Google Scholar] [CrossRef] [PubMed]
  95. Österlund, M.K. Underlying mechanisms mediating the antidepressant effects of estrogens. Biochem. Biophys. Acta 2010, 1800, 1136–1144. [Google Scholar] [CrossRef]
  96. McQueen, J.K.; Wilson, H.; Fink, G. Estradiol-17β increase serotonin transporter (SERT) mRNA levels and density of SERT-binding sites in female rat brain. Brain Res. Mol. Brain Res. 1997, 45, 13–23. [Google Scholar] [CrossRef] [PubMed]
  97. Sanchez, M.G.; Estrada-Camarena, E.; Belanger, N.; Morissette, M.; Di Paolo, T. Estradiol modulation of cortical, striatal and raphe nucleus 5-HT1A and 5-HT2A receptors of female hemiparkinsonian monkeys after long-term ovariectomy. Neuropharmacology 2011, 60, 642–652. [Google Scholar] [CrossRef]
  98. Sanchez, M.G.; Morissette, M.; Di Paolo, T. Osteradiol modulation of serotonin reuptake transporter and serotonin metabolism in the brain of monkeys. J. Neuroendocrinol. 2013, 25, 560–569. [Google Scholar] [CrossRef]
  99. Nakashima, A.; Ota, A.; Sabban, E.L. Interaction between Egr1 and AP1 factors in regulation of tyrosine hydroxylase transcription. Brain Res. Mol. Brain Res. 2003, 112, 61–69. [Google Scholar] [CrossRef]
  100. Lewis-Tuffin, L.J.; Quinn, P.G.; Chikaraishi, D.M. Tyrosine hydroxylase transcription depends primarily on cAMP response element activity, regardless of the type of inducing stimulus. Mol. Cell. Neurosci. 2004, 25, 536–547. [Google Scholar] [CrossRef]
  101. Maharjan, S.; Serova, L.; Sabban, E.L. Transcriptional regulation of tyrosine hydroxylase by estrogen: Opposite effects with estrogen receptor a and b and interactions with cyclic AMP. J. Neurochem. 2005, 93, 1503–1514. [Google Scholar] [CrossRef]
  102. Liaw, J.J.; He, J.R.; Hartman, R.D.; Barraclough, C.A. Changes in tyrosine hydroxylase mRNA levels in medullary A and A2 neurons and locus coeruleus following castration and estrogen replacement in rats. Brain Res. Mol. Brain Res. 1992, 13, 231–238. [Google Scholar] [CrossRef] [PubMed]
  103. Serova, L.; Rivkin, M.; Nakashima, A.; Sabban, E.L. Estradiol stimulates gene expression of norepinephrine biosynthetic enzymes in rat locus coeruleus. Neuroendocrinology 2002, 75, 193–200. [Google Scholar] [CrossRef] [PubMed]
  104. Serova, L.I.; Maharjan, S.; Huang, A.; Sun, D.; Kaley, G.; Sabban, E.L. Response of tyrosine hydroxylase and GTP cyclohydrolase I gene expression to estrogen in the brain catecholaminergic regions varies with mode of administration. Brain Res. 2004, 1015, 1–8. [Google Scholar] [CrossRef]
  105. Thanky, N.R.; Son, J.H.; Herbison, A.E. Sex differences in the regulation of tyrosine hydroxylase gene transcription by estrogen in locus coeruleus of TH9-LacZ transgenic mice. Brain Res. Mol. Brain Res. 2002, 104, 220–226. [Google Scholar] [CrossRef] [PubMed]
  106. Maharjan, S.; Serova, L.I.; Sabban, E.L. Membrane-initiated estradiol signaling increases tyrosine hydroxylase promoter activity with ERα in PC12 cells. J. Neurochem. 2010, 112, 42–55. [Google Scholar] [CrossRef]
  107. Yanagihara, N.; Liu, M.; Toyohira, Y.; Tsutsui, M.; Ueno, S.; Shinohara, Y.; Takahashi, K.; Tanaka, K. Stimulation of catecholamine synthesis through unique estrogen receptors in the bovine adrenomedullary plasma membrane by 17β-estradiol. Biochem. Biophys. Res. Commun. 2006, 339, 548–553. [Google Scholar] [CrossRef]
  108. Inagaki, H.; Toyohira, Y.; Takahashi, K.; Ueno, S.; Kawagoe, T.; Tsutsui, M.; Hachisuga, T.; Yanagisawa, N. Effects of selective estrogen receptor modulators on plasma membrane estrogen receptors and catecholamine synthesis and secretion in cultured bovine adrenal medullary cells. J. Pharmacol. Sci. 2014, 124, 66–75. [Google Scholar] [CrossRef]
  109. Epperson, C.N.; Kim, D.R.; Bale, T.L. Estradiol modulation of monoamine metabolism one possible mechanism underlying sex differences in risk for depression and dementia. JAMA Psychiatry 2014, 71, 869–870. [Google Scholar] [CrossRef]
  110. Sacher, J.; Rabiner, E.A.; Clark, M.; Rusjan, P.; Soliman, A.; Boskovic, R.; Kish, S.J.; Wilson, A.A.; Houle, S.; Meyer, J.H. Dynamic, adaptive changes in MAO-A binding after alterations in substrate availability: An in vivo [11C]-harmine position emission tomography study. J. Cereb. Blood Flow. Metab. 2012, 32, 443–446. [Google Scholar] [CrossRef]
  111. Rekkas, P.V.; Wilson, A.A.; Lee, V.W.H.; Yogalingam, P.; Sacher, J.; Rusjan, P.; Houle, S.; Stewart, D.E.; Kolla, N.J.; Kish, S.; et al. Greater monoamine oxidase A binding in perimenopausal age as measured with carbon 11-labeled harmine position emission tomography. JAMA Psychiatry 2014, 17, 873–879. [Google Scholar] [CrossRef] [PubMed]
  112. Holschneider, D.P.; Kumazawa, T.; Chen, K.; Shih, J.C. Tissue-specific effects of estrogen on monoamine oxidase A and B in the rat. Life Sci. 1998, 63, 155–160. [Google Scholar] [CrossRef] [PubMed]
  113. Bethea, C.L.; Lu, N.Z.; Gundlah, C.; Streicher, J.M. Diverse actions of ovarian steroids in the serotonin neural system. Front. Neuroendocrinol. 2002, 23, 41–100. [Google Scholar] [CrossRef]
  114. Wu, J.B.; Chen, K.; Li, Y.; Lau, Y.F.C.; Shih, J.C. Regulation of monoamine oxidase A by the SRY gene on the Y chromosome. FASEB J. 2009, 23, 4029–4038. [Google Scholar] [CrossRef] [PubMed]
  115. Shih, J.C.; Wu, J.; Chen, K. Transcriptional regulation and multiple functions of MAO genes. J. Neural Transm. 2011, 118, 979–986. [Google Scholar] [CrossRef]
  116. Ou, X.M.; Chen, K.; Shih, J.C. Glucocorticoid and androgen activation of monoamine oxidase A is regulated by R1 and Sp1. J. Biol. Chem. 2006, 281, 21512–21525. [Google Scholar] [CrossRef]
  117. Kranz, G.S.; Spies, M.; Vraka, C.; Kaufmann, U.; Klebermass, E.M.; Handschuh, P.A.; Ozenil, M.; Murgaš, M.; Pichler, V.; Rischka, L.; et al. High-dose testosterone treatment reduces monoamine oxidase A levels in human brain: A preliminary report. Psychoneuroendocrinology 2021, 133, 105381. [Google Scholar] [CrossRef]
  118. Xie, T.; Ho, S.L.; Ramsden, D. Characterization and implications of estrogenic down-regulation of human catechol-O-methyltransferase gene transcription. Mol. Pharmacol. 1999, 56, 31–38. [Google Scholar] [CrossRef]
  119. Jiang, H.; Xie, T.; Ramsden, D.B.; Ho, S.L. Human catechol-O-methyltransferase downregulation by estradiol. Neuropharmacology 2003, 45, 1011–1018. [Google Scholar] [CrossRef]
  120. Yao, J.; Li, Y.; Chang, M.; Wu, H.; Yang, X.; Goodman, J.E.; Liu, X.; Liu, H.; Mesecar, A.D.; Van Breemen, R.B.; et al. Catechol estrogen 4-hydroxyequilenin is a substrate and an inhibitor of catechol-O-methyltransferase. Chem. Res. Toxicol. 2003, 16, 668–675. [Google Scholar] [CrossRef]
  121. Patel, P.D.; Bochar, D.; Tumer, D.L.; Meng, F.; Moeller, H.M.; Pontrello, C.G. Regulation of tryptophan hydroxylase-2 gene expression by a bipartite RE-1 silencer of transcription/neuron-restrictive silencing factor (REST/NRSF) binding motif. J. Biol. Chem. 2007, 282, 26717–26724. [Google Scholar] [CrossRef] [PubMed]
  122. Clark, J.A.; Flick, R.B.; Pai, L.Y.; Szalayova, I.; Key, S.; Conley, R.K.; Deutch, A.Y.; Hutson, P.H.; Mazey, E. Glucocorticoid modulation of tryptophan hydroxylase-2 protein in raphe nuclei and 6-hydroxytryptophan concentrations in frontal cortex of C56/B16 mice. Mol. Psychiatry 2008, 13, 498–506. [Google Scholar] [CrossRef]
  123. Malek, Z.S.; Sage, D.; Pevet, P.; Raison, S. Daily rhythm of tryptophan hydroxylase-2 messenger ribonucleic acid within raphe neurons is induced by corticoid daily surge and modulated by enhanced locomotor activity. Endocrinology 2007, 148, 5165–5172. [Google Scholar] [CrossRef]
  124. Sabban, E.; Nankova, B.B.; Serova, L.I.; Kvetnansky, R.; Liu, X. Molecular regulation of gene expression of catecholamine biosynthetic enzymes by stress: Sympathetic ganglia versus adrenal medulla. Ann. N. Y. Acad. Sci. 2004, 1018, 370–377. [Google Scholar] [CrossRef] [PubMed]
  125. Rani, C.S.S.; Elango, N.; Wang, S.S.; Kobayashi, K.; Strong, R. Identification of an activator protein-1-like sequence as the glucocorticoid response element in the rat tyrosine hydroxylase. Mol. Pharmacol. 2009, 75, 589–598. [Google Scholar] [CrossRef]
  126. Grunewald, M.; Johnson, S.; Lu, D.; Wang, Z.; Lomberk, G.; Albert, P.R.; Stockmeier, C.A.; Meyer, J.H.; Urrutia, R.; Miczek, K.A.; et al. Mechanistic role for a novel glucocorticoid-KLF11(TIEG2) protein pathway in stress-induced monoamine oxidase A expression. J. Biol. Chem. 2012, 287, 24195–24205. [Google Scholar] [CrossRef]
  127. Walf, A.A.; Frye, C.A. Administration of estrogen receptor beta-specific selective estrogen receptor modulators to the hippocampus decrease anxiety and depressive behavior of ovariectomized rats. Pharmacol. Biochem. Behav. 2007, 86, 407–414. [Google Scholar] [CrossRef] [PubMed]
  128. Rietjens, I.M.C.M.; Louisse, J.; Beekmann, K. The potential health effects of dietary phytoestrogens. Br. J. Pharmacol. 2017, 174, 1263–1280. [Google Scholar] [CrossRef]
  129. Strauss, L.; Santti, R.; Saarinen, N.; Streng, T.; Joshi, S.; Mäkelä, S. Dietary phytoestrogens and their role in hormonally dependent disease. Toxicol. Lett. 1998, 102–103, 349–354. [Google Scholar] [CrossRef]
  130. Tice, J.A.; Ettinger, B.; Ensrud, K.; Wallace, R.; Blackwell, T.; Cummings, S.R. Phytoestrogen supplements for the treatment of hot flashes: The isoflavone clover extract (ICE) study: A randomized controlled trial. JAMA 2003, 290, 207–214. [Google Scholar] [CrossRef]
  131. Chen, L.R.; Chen, K.H. Utilization of isoflavones in soybeans for women with menopausal syndrome: An overview. Int. J. Mol. Sci. 2021, 22, 3212. [Google Scholar] [CrossRef] [PubMed]
  132. Kuiper, G.G.J.M.; Lemmen, J.B.; Carlsson, B.; Corton, J.C.; Safe, S.H.; van der Saag, P.T.; van der Burg, B.; Gustafsson, J.A. Interaction of estrogenic chemicals and phytoestrogens with estrogen receptor-β. Endocrinology 1998, 139, 4252–4263. [Google Scholar] [CrossRef]
  133. Kostelac, D.; Rechkemmer, G.; Briviba, K. Phytoestrogens modulate binding response to estrogen receptors a and b to the estrogen response element. J. Agric. Food Chem. 2003, 51, 7632–7635. [Google Scholar] [CrossRef] [PubMed]
  134. Van Duursen, M.B.M. Modulation of estrogen synthesis and metabolism by phytoestrogens in vitro and the implications for women’s health. Toxicol. Res. 2017, 6, 772–794. [Google Scholar] [CrossRef] [PubMed]
  135. De La Cruz, C.P.; Revilla, E.; Venero, J.L.; Ayala, A.; Cano, J.; Machado, A. Oxidative inactivation of tyrosine hydroxylase in substantia nigra of aged rat. Free Radic. Biol. Med. 1996, 20, 53–61. [Google Scholar] [CrossRef]
  136. Hussain, A.M.; Mitra, A.K. Effect of aging on tryptophan hydroxylase in rat brain: Implications on serotonin level. Drug Metab. Dispos. 2000, 28, 1038–1042. [Google Scholar] [CrossRef]
  137. Sarubbo, E.; Ramis, M.R.; Kienzer, C.; Aparicio, S.; Esteban, S.; Miralles, A.; Moranta, D. Chronic silymarin, quercetin and naringenin treatment increase monoamines synthesis and hippocampal Sirt1 levels improving cognition in aged rats. J. Neuroimmme. Pharmacol. 2018, 13, 24–38. [Google Scholar] [CrossRef]
  138. Ramis, M.R.; Sarubbo, F.; Tejada, S.; Jimenez, M.; Esteban, S.; Miralles, A.; Moranta, D. Chronic polyphenon-60 or catechin treatments increase brain monoamines syntheses and hippocampal SIRT1 levels improving cognition on aged rats. Nutrients 2020, 12, 326. [Google Scholar] [CrossRef]
  139. Sarubbo, E.; Ramis, M.R.; Aparicio, S.; Ruiz, L.; Esteban, S.; Miralles, A.; Moranta, D. Improving effects of chronic resveratrol treatment on central monoamine synthesis and cognition in aged rats. Age 2015, 37, 9777. [Google Scholar] [CrossRef]
  140. Lee, B.; Choi, G.M.; Sur, B. Silibinin prevents depression-like behaviors in a single prolonged stress rat model: The possible role of serotonin. BMC Complement. Med. Ther. 2020, 20, 70. [Google Scholar] [CrossRef]
  141. Lee, B.; Sur, B.; Lee, H.; Oh, S. Korean red ginseng posttraumatic stress disorder-triggered depression-like behaviors in rats via activation of the serotonergic system. J. Ginseng Res. 2020, 44, 644–654. [Google Scholar] [CrossRef] [PubMed]
  142. Lee, B.; Choi, G.M.; Sur, B. Anti-depressant-like effects of 142 in animal model of post-traumatic stress disorder. Clin. J. Interg. Med. 2021, 27, 39–46. [Google Scholar] [CrossRef]
  143. Vauzour, D.; Rendeiro, C.; D’Amatao, A.; Waffo-Téguo, P.; Richard, T.; Mérillon, J.M.; Pontifex, M.G.; Connell, E.; Müller, M.; Butler, L.T.; et al. Anthrocyanins promote learning through modulation of synaptic plasticity related proteins in animal model of ageing. Antioxidants 2021, 10, 1235. [Google Scholar] [CrossRef]
  144. Zhang, X.; Xiong, J.; Liu, S.; Wang, L.; Huang, J.; Liu, L.; Yang, J.; Zhang, G.; Guo, K.; Zhang, Z.; et al. Puerarin protects dopaminergic neurons in Parkinson’s disease models. Neuroscience 2014, 280, 88–98. [Google Scholar] [CrossRef] [PubMed]
  145. Sonia Angeline, M.; Sarkar, A.; Anand, K.; Ambasta, R.K.; Kumar, P. Sesamol and naringenin reverse the effect of rotenone-induced PD rat model. Neuroscience 2013, 254, 379–394. [Google Scholar] [CrossRef]
  146. Singh, S.S.; Rai, S.N.; Birla, H.; Zahra, W.; Kumar, G.; Gedda, M.R.; Tiwari, N.; Patnaik, R.; Singh, R.K.; Singh, S.P. Effect of chlorogenic acid supplementation in MPTP-intoxicated mouse. Front. Pharmacol. 2018, 9, 757. [Google Scholar] [CrossRef]
  147. Rose, K.; Parmar, M.; Cavanaugh, J.E. Dietary supplementation with resveratrol protects against striatal dopaminergic deficits produced by in utero LPS exposure. Brain Res. 2014, 1573, 37–43. [Google Scholar] [CrossRef]
  148. Zhang, G.; Yang, G.; Liu, J. Phloretin attenuates behavior deficits and neuroinflammatory response in MPTP induced Parkinson’s disease in mice. Life Sci. 2019, 232, 116600. [Google Scholar] [CrossRef]
  149. Fatima, A.; Rahul; Siddique, Y.H. Role of tangeritin against cognitive impairments in transgenic Drosophila model of Parkinson’s disease. Neurosci. Lett. 2019, 705, 112–117. [Google Scholar] [CrossRef]
  150. Saied, N.M.; Georgy, G.S.; Hussien, R.M.; Hassan, W.A. Neuromodulatory effect of curcumin on catecholamine systems and inflammatory cytokines in ovariectomized female rats. Clin. Exp. Pharmacol. Physiol. 2021, 48, 337–346. [Google Scholar] [CrossRef]
  151. Hack, W.; Gladen-Kolarsky, N.; Chatterjee, S.; Liang, Q.; Maitra, U.; Giesla, L.; Gray, N.E. Gardenin A treatment attenuates inflammatory markers, synuclein pathology and deficits in tyrosine hydroxylase expression and improves cognitive and motor function in A53T-a-syn mice. Biomed. Pharmacother. 2024, 173, 116370. [Google Scholar] [CrossRef]
  152. Liu, M.; Yanagihara, N.; Toyohira, Y.; Tsutsui, M.; Ueno, S.; Shinohara, Y. Dual effects of daidzen, and soy isoflavone, on catecholamine synthesis in cultured bovine adrenal medullary cells. Endocrinology 2007, 148, 5348–5354. [Google Scholar] [CrossRef] [PubMed]
  153. Zhang, H.; Yanagihara, N.; Toyohira, Y.; Takahashi, K.; Inagaki, H.; Satoh, N.; Li, X.; Goa, X.; Tsutsui, M.; Takahaishi, K. Stimulatory effect of nobiletin, a citrus polymethoxy flavone, on catecholamine synthesis through Ser19 and Ser40 phosphorylation of tyrosine hydroxylase in cultured bovine adrenal medullary cells. Naunyn-Schmiedebergs Arch. Pharmacol. 2014, 387, 15–22. [Google Scholar] [CrossRef] [PubMed]
  154. Vina, D.; Serra, S.; Lamela, M.; Delogu, G. Herbal natural products as a source of monoamine oxidase inhibitors: A review. Curr. Top. Med. Chem. 2012, 12, 2131–2144. [Google Scholar] [CrossRef]
  155. Chimenti, F.; Cottiglia, F.; Bonsignore, L.; Casu, L.; Casu, M.; Floris, C.; Secci, D.; Bolasco, A.; Chimenti, P.; Granese, A.; et al. Quercetin as the active principle of Hypericum hircinum exerts a selective inhibitory activity against MAO-A: Extraction, biological analysis, and computational study. J. Nat. Prod. 2006, 69, 945–949. [Google Scholar] [CrossRef]
  156. Carradori, S.; Gidaro, M.C.; Petzer, A.; Costa, G.; Guglielmi, P.; Chimenti, P.; Alcaro, S.; Petzer, J.P. Inhibition of human monoamine oxidase: Biological and molecular modeling studies on selected natural flavonoids. J. Agric. Food Chem. 2016, 64, 9004–9011. [Google Scholar] [CrossRef] [PubMed]
  157. Adeoluwa, O.A.; Eduviere, A.T.; Adeoluwa, G.O.; Otomewo, L.O.; Adeniyi, F.R. The monoaminergic pathways are involved in the antidepressant-like effect of quercetin. Naunyn Schmiedebergs Arch. Pharmacol. 2024, 397, 2497–2506. [Google Scholar] [CrossRef]
  158. Bandaruk, Y.; Mukai, R.; Kawamura, T.; Nemoto, H.; Terao, J. Evaluation of the inhibitory effects of quercetin-related flavonoids and tea catechins on the monoamine oxidase-A reaction in mouse brain mitochondria. J. Agric. Food Chem. 2012, 60, 10270–10277. [Google Scholar] [CrossRef]
  159. Lee, H.W.; Ryu, H.W.; Kang, M.G.; Park, D.; Oh, S.R.; Kim, H. Selective inhibition of monoamine oxidase A by purpurin, an anthraquinone. Bioorg. Med. Chem. Lett. 2017, 27, 1136–1140. [Google Scholar] [CrossRef]
  160. Gidaro, M.C.; Astorino, C.; Petzer, A.; Carradori, S.; Alcaro, F.; Costa, G.; Artese, A.; Rafele, G.; Russo, F.M.; Petzer, J.P.; et al. Kaempferol as selective MAO-A inhibitor: Analytical detection in Calabrian red wines, biological and molecular modeling studies. J. Agric. Food Chem. 2016, 64, 1394–1400. [Google Scholar] [CrossRef]
  161. Olayinka, J.N.; Akawa, O.B.; Ogbu, E.K.; Eduviere, A.T.; Ozolua, R.I.; Soliman, M. Apigenin attenuates depressive-like behavior via modulating monoamine oxidase A enzyme activity in chronically stressed mice. Curr. Res. Pharnacol. Drug Discov. 2023, 5, 100161. [Google Scholar] [CrossRef]
  162. Guo, B.; Zheng, C.; Cai, W.; Cheng, J.; Wang, H.; Li, H.; Sun, Y.; Cui, W.; Wang, Y.; Han, Y.; et al. Multifunction of chrysin in Parkinson’s model: Antineuronal apoptosis, neuroprotection via activation of MEF2D, and inhibition of monoamine oxidase-B. J. Agric. Food Chem. 2016, 64, 5324–5333. [Google Scholar] [CrossRef]
  163. Yanez, M.; Fraiz, N.; Cano, E.; Orallo, F. (-)-Trans-ε-viniferin, a polyphenol present in wines, is an inhibitor of noradrenaline and 5-hydroxytryptamine uptake and of monoamine oxidase activity. Eur. J. Pharmacol. 2006, 542, 54–60. [Google Scholar] [CrossRef]
  164. Carradori, S.; D’Ascenzio, M.; Chimenti, P.; Secchi, D.; Bolasco, A. Selective MAO-B inhibitors: A lesson from natural products. Mol. Divers. 2014, 18, 219–243. [Google Scholar] [CrossRef]
  165. Minders, C.; Petzer, J.P.; Petzer, A.; Lourens, A.C.U. Monoamine oxidase inhibitory activities of heterocyclic chalcones. Bioorg. Med. Chem. Lett. 2015, 25, 5270–5276. [Google Scholar] [CrossRef] [PubMed]
  166. Sashidhara, K.V.; Modukuri, R.K.; Singh, S.; Rao, K.B.; Teja, G.A.; Gupta, S.; Shukla, S. Design and synthesis of new series of coumarin-aminopyran derivatives possessing potential anti-depressant-like activity. Bioorg. Med. Chem. Lett. 2015, 25, 337–341. [Google Scholar] [CrossRef] [PubMed]
  167. Bergström, M.; Westerberg, G.; Nemeth, G.; Traut, M.; Gross, G.; Greger, G.; Müller-Peltzer, H.; Safer, A.; Eckernäs, S.A.; Grahnér, A.; et al. MAO-A inhibition in brain after dosing with esuprone, moclobemide and placebo in healthy volunteers: In vivo studies with position emission tomography. Eur. J. Clin. Pharmacol. 1997, 52, 121–129. [Google Scholar] [CrossRef]
  168. Matos, M.J.; Teran, C.; Perez-Castillo, Y.; Uriarte, E.; Santana, L.; Vina, D. Synthesis and study of a series of 3-arylcoumarins as potent and selective monoamine oxidase B inhibitors. J. Med. Chem. 2011, 54, 7127–7137. [Google Scholar] [CrossRef]
  169. Abdelhafez, O.; Amin, K.M.; Ali, H.I.; Abdalla, M.; Batran, R.Z. Synthesis of new 7-oxycoumarin derivatives as potent and selective monoamine oxidase A inhibitors. J. Med. Chem. 2012, 55, 10424–10436. [Google Scholar] [CrossRef]
  170. Bindra, S.; Datta, A.; Yasin, H.K.A.; Hanttu, A.; Pollesello, P. Recent progress in synthetic and natural catechol-O-methyltransferase inhibitors for neurological disorders. ACS Omega 2024, 9, 44005–44018. [Google Scholar] [CrossRef]
  171. Hitge, R.; Smit, S.; Petzer, A.; Petzer, J. Evaluation of nitrocatechol chalcone and pyrazoline derivatives as inhibitors of catechol-O-methyltransferase and monoamine oxidase. Bioorg. Med. Chem. Lett. 2020, 30, 127188. [Google Scholar] [CrossRef] [PubMed]
  172. Engelbrecht, I.; Petzer, J.P.; Petzer, A. Evaluation of selective natural compounds as dual inhibitors of catechol-O-methyltransferase and monoamine oxidase. Cent. Nerv. Syst. Agents Med. Chem. 2019, 19, 133–145. [Google Scholar] [CrossRef] [PubMed]
  173. Kadowaki, W.; Miyata, R.; Fuzinami, M.; Sato, Y.; Kumazawa, S. Catechol-O-methyltransferase inhibitors from Calendula officinalis leaf. Molecules 2023, 28, 1333. [Google Scholar] [CrossRef]
  174. Chen, D.; Wang, C.Y.; Lambert, J.D.; Ai, N.; Welsh, W.J.; Yang, C.S. Inhibition of human liver catechol-O-methyltransferase by tea catechins and their metabolites: Structure-activity relationship and molecular-modeling studies. Biochem. Pharmacol. 2005, 69, 1523–1531. [Google Scholar] [CrossRef]
  175. Tammen, S.A.; Friso, S.; Choi, S.W. Epigenetics: The link between nature and nurture. Mol. Aspect Med. 2013, 34, 753–764. [Google Scholar] [CrossRef] [PubMed]
  176. Martin, S.L.; Hardy, T.M.; Tollefsbol, T.O. Medical chemistry of epigenetic diet and caloric restriction. Curr. Med. Chem. 2013, 20, 4050–4059. [Google Scholar] [CrossRef] [PubMed]
  177. Wachs, T.D.; Georgieff, M.; Cusick, S.; McEwen, B.S. Issues in the timing of integrated early interventions: Contributions from nutrition, neuroscience, and psychological research. Ann. N. Y. Acad. Sci. 2014, 1308, 89–106. [Google Scholar] [CrossRef]
  178. Tobi, E.W.; Goeman, J.J.; Monajemi, R.; Gu, H.; Putter, H.; Zhang, Y.; Slieker, R.C.; Stok, A.P.; Thijssen, P.E.; Müller, F.; et al. DNA methylation signatures link prenatal famine exposure to growth and metabolism. Nat. Commun. 2014, 5, 5592. [Google Scholar] [CrossRef]
  179. de la Garza, A.L.; Garza-Cuellar, M.A.; Silva-Hernandez, I.A.; Cardenas-Perez, R.E.; Reyes-Castro, L.A.; Zambrano, E.; Gonzalez-Hernandez, B.; Garza-Ocañas, L.; Fuentes-Mera, L.; Camacho, A. Maternal flavonoids intake reverts depression-like behaviour in rat female offspring. Nutrients 2019, 11, 572. [Google Scholar] [CrossRef]
  180. Booij, L.; Tremblay, R.E.; Szyf, M.; Benkelfat, C. Genetic and early environmental influences on the serotonin system: Consequences for brain development and risk for psychopathology. J. Psychiatry Neurosci. 2015, 40, 5–18. [Google Scholar] [CrossRef]
  181. Lambe, E.K.; Krimer, L.S.; Goldman-Rakic, P.S. Differential postnatal development of catecholamine and serotonin inputs to identified neurons in prefrontal cortex of rhesus monkey. J. Neurosci. 2000, 20, 8780–8787. [Google Scholar] [CrossRef]
  182. Suri, D.; Teixeira, C.M.; Cagliostro, M.K.C.; Mahadevia, D.; Ansorge, M.S. Monoamine-sensitive developmental periods impacting adult emotional and cognitive behaviors. Neuropsychopharmacology 2015, 40, 88–112. [Google Scholar] [CrossRef]
  183. Cheng, Z.; Su, J.; Zhang, K.; Jaing, H.; Li, B. Epigenetic mechanism of early life stress-induced depression: Focus on the neurotransmitter systems. Front. Cell Dev. Biol. 2022, 10, 929732. [Google Scholar] [CrossRef] [PubMed]
  184. Hellstrom, I.C.; Dhir, S.K.; Diorio, J.C.; Meaney, M.J. Maternal licking regulates hippocampal glucocorticoid receptor transcription through a thyroid-hormone-serotonin-NGFI-A signalling cascade. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2012, 367, 2495–2510. [Google Scholar] [CrossRef] [PubMed]
  185. Mazer, C.; Muneyyirci, J.; Taheny, K.; Raio, N.; Borella, A.; Whitaker-Azmitia, P. Serotonin depletion during synaptogenesis leads to decreased synaptic density and learning deficits in the adult rat: A possible model of neurodevelopmental disorders with cognitive deficits. Brain Res. 1997, 760, 68–73. [Google Scholar] [CrossRef] [PubMed]
  186. Cases, O.; Vitalis, T.; Seif, I.; De Maeyer, E.; Sotelo, C.; Gasper, P. Lack of barrels in the somatosensory cortex of monoamine oxidase A-deficit mice: Role of a serotonin excess during the critical period. Neuron 1996, 16, 297–307. [Google Scholar] [CrossRef] [PubMed]
  187. Chen, K.; Cases, O.; Rebrin, I.; Wu, W.; Gallaher, T.K.; Shih, J.C. Forebrain-specific expression of monoamine oxidase A reduces neurotransmitter levels, restores the brain structure, and rescues aggressive behavior in monoamine oxidase A-deficit mice. J. Biol. Chem. 2007, 282, 115–123. [Google Scholar] [CrossRef]
  188. Srinivas, T.; Mathias, C.; Oliveira-Mateos, C.; Guil, S. Roles of lncRNAs in brain development and pathogenesis: Emerging therapeutic opportunities. Mol. Ther. 2023, 31, 1550–1561. [Google Scholar] [CrossRef]
  189. Jovcevska, I.; Paska, A.V. Neuroepigenetics of psychiatric disorders: Focus on lncRNA. Neurochem. Int. 2021, 149, 1054140. [Google Scholar] [CrossRef]
  190. Zhong, X.L.; Du, Y.; Chen, L.; Cheng, Y. The emerging role of long noncoding RNA in depression and its implications in diagnostics and therapeutic responses. J. Psychiatr. Res. 2023, 164, 251–258. [Google Scholar] [CrossRef]
  191. Scott, K.M.; McLaughlin, K.A.; Smith, D.A.R.; Ellis, P.M. Childhood maltreatment and DSM-IV adult mental disorders: Comparison of prospective and retrospective findings. Br. J. Psychiatry 2012, 200, 469–475. [Google Scholar] [CrossRef] [PubMed]
  192. Torres-Berrio, A.; Issler, O.; Parise, E.M.; Nestler, E.J. Unveiling the epigenetic landscape of depression: Focus on early life stress. Dialogues Clin. Neurosci. 2019, 21, 341–357. [Google Scholar] [CrossRef]
  193. Konig, S.M.; Kessler, C.L.; Canli, T.; Duman, E.A.; Ada, E.K.; Zinbarg, R.; Craske, M.G.; Stephens, J.E.; Vrshek-Schallhorm, S. Early-life adversity severity, timing, and context type are associated with SLC6A4 methylation in emerging adults: Results from a prospective cohort study. Psychoneuroendocrinolgy 2024, 170, 107181. [Google Scholar] [CrossRef]
  194. Wang, D.; Szyf, M.; Benkelfat, C.; Provencal, N.; Turecki, G.; Caramaschi, D.; Cote, S.M.; Vitaro, F.; Tremblay, R.E.; Booij, L. Peripheral SLC6A4 DNA methylation is associated with in vivo measures of human brain serotonin synthesis and childhood physical aggression. PLoS ONE 2012, 7, e39501. [Google Scholar] [CrossRef] [PubMed]
  195. Carrard, A.; Salzmann, A.; Malafosse, A.; Karege, F. Increased DNA methylation status of the serotonin receptor 5HTR1A gene promoter in schizophrenia and bipolar disorder. J. Affect. Disord. 2011, 132, 450–453. [Google Scholar] [CrossRef]
  196. Quellet-Morin, I.; Wong, C.C.Y.; Danese, A.; Parieante, C.M.; Papadopoulos, A.S.; Mill, J.; Arsenault, L. Increased serotonin transporter gene (SERT) DNA methylation is associated with bullying victimization and blunt cortisol response to stress in childhood: A longitudinal study of discordant monozygotic twins. Psychol. Med. 2013, 13, 1813–1823. [Google Scholar] [CrossRef]
  197. Checknita, D.; Tihonen, J.; Hodgins, S.; Nilsson, K. Association of age, sex, sexual abuse, and gene type with monoamine oxidase gene a gene methylation. J. Neural Transm. 2021, 128, 1721–1739. [Google Scholar] [CrossRef] [PubMed]
  198. Rahman, I.; Chung, S. Dietary polyphenols, deacetylases and chromatin remodeling in inflammation. J. Nutrigenet. Neurogenomics 2010, 3, 220–230. [Google Scholar] [CrossRef]
  199. Schneider-Stock, R.; Ghantous, A.; Bajboul, K.; Saikali, M.; Darwiche, N. Epigenetic mechanisms of plant-derived anticancer drugs. Front. Biosci. 2012, 17, 129–173. [Google Scholar] [CrossRef]
  200. Remely, M.; Lovrecic, L.; de la Garza, A.L.; Magliore, L.; Peterlin, B.; Milagro, F.I.; Martinez, A.J.; Haslberger, A.G. Therapeutic perspectives of epigenetically active nutrients. Br. J. Pharmacol. 2015, 172, 2756–2768. [Google Scholar] [CrossRef]
  201. Choi, S.W.; Friso, S. Epigenetics: A new bridge between nutrition and health. Adv. Nutr. 2010, 1, 8–16. [Google Scholar] [CrossRef] [PubMed]
  202. Ayissi, V.B.; Ebrahimi, A.; Schluesenner, H. Epigenetic effects of natural polyphenols: A focus on SIRT1-mediated mechanisms. Mol. Nutr. Food Res. 2014, 58, 22–32. [Google Scholar] [CrossRef] [PubMed]
  203. de Boer, V.C.; de Goffau, M.C.; Arts, I.C.W.; Hollman, P.C.H.; Keijer, J. SIRT1 stimulation by polyphenols is affected by their stability and metabolism. Mech. Ageing Dev. 2006, 127, 618–627. [Google Scholar] [CrossRef] [PubMed]
  204. Patil, S.P.; Tran, N.; Geekiyanage, H.; Liu, L.; Chan, C. Curcumin-induced upregulation of the anti-tau cochaperone BAG2 in primary rat cortical neurons. Neurosci. Lett. 2013, 554, 121–125. [Google Scholar] [CrossRef]
  205. Taha, M.; Eldemerdash, O.M.; Elshaffei, I.M.; Yousef, E.M.; Soliman, A.S.; Senousy, M.A. Apigenin attenuates hippocampal microglial activation and restores cognitive function in methotrexate-treated rats: Targeting the miR-15a/Rock-1/ERK1/2 pathway. Mol. Neurobiol. 2023, 60, 3770–3787. [Google Scholar] [CrossRef]
  206. Habibi, P.; Babri, S.; Ahmadiasl, N.; Yousefi, H. Effects of genistein and swimming exercise on spatial memory and expression of microRNA 132, BDNF, and IGF-1 genes in the hippocampus of ovariectomized rats. Iran. J. Basic. Med. Sci. 2017, 20, 856–862. [Google Scholar] [CrossRef]
  207. Gao, Y.; Zhang, N.; Lv, C.; Li, X.; Li, W. lncRNA SNGH1 knockdown alleviates amyloid-β-induced neuronal injury by regulating ZNF27 vi sponging miR-361-3- in Alzheimer’s disease. J. Alzheimers Dis. 2020, 77, 85–98. [Google Scholar] [CrossRef] [PubMed]
  208. Xia, D.; Sui, R.; Zhang, Z. Administration of resveratrol improved Parkinson’s disease-like phenotype by suppressing apoptosis of neurons via modulating the MALTAT1/miR-129/SNCA signaling pathway. J. Cell. Biochem. 2019, 120, 4942–4951. [Google Scholar] [CrossRef]
  209. Li, X.; Su, Y.; Li, N.; Zhang, F.R.; Zhang, N. Berberine attenuates MPP+-induced neuronal injury by regulating LINGC00943/miR-142-5p/KPNA4/NF-κB pathway in SK-N-SH cells. Neurochem. Res. 2021, 46, 3286–3300. [Google Scholar] [CrossRef]
  210. Zhang, S.; Wang, H.; Wang, J.; Jin, W.; Yan, X.; Chen, X.; Wang, D.; Zhao, D.; Wang, Y.; Cong, D.; et al. Ginsenoside Rf inhibits human tau proteotoxicity and specific LncRNA, miRNA and mRNA expression changes in Caenorhabditis elegans model of tauopathy. Eur. J. Pharmacol. 2022, 922, 174887. [Google Scholar] [CrossRef]
  211. Hardy, T.M.; Tollefsbol, T. Epigenetic diet: Impact on the epigenome and cancer. Epigenomics 2011, 3, 503–518. [Google Scholar] [CrossRef]
  212. Prasanth, M.I.; Sivamaruthi, B.S.; Cheong, C.S.Y.; Verma, K.; Tencomnao, T.; Brimson, J.M.; Prasansuklab, A. Role of epigenetic modulation in neurodegenerative diseases: Implications of phytochemical intervention. Antioxidants 2024, 13, 606. [Google Scholar] [CrossRef] [PubMed]
  213. Qin, W.; Zhu, W.; Shi, H.; Hewtt, J.E.; Rohlen, R.L.; MacDonald, R.S.; Rottinghaus, G.E.; Chen, Y.C.; Sauter, E.R. Soy isoflavones have an antiestrogenic effect and alter mammary promoter hypermethylation in healthy premenopausal women. Nutr. Cancer 2009, 61, 238–244. [Google Scholar] [CrossRef] [PubMed]
  214. Atteritano, M.; Mazzaferro, S.; Bitto, A.; Cannata, M.L.; D'Anna, R.; Squadrito, F.; Macrì, I.; Frisina, A.; Frisina, N.; Bagnato, G. Genistein effects on quality of life and depression symptoms in osteopenic postmenopausal women: A 2-year randomized, double-blind, controlled study. Osteoporos. Int. 2014, 25, 1123–1129. [Google Scholar] [CrossRef] [PubMed]
  215. Behl, T.; Rana, T.; Sehgal, A.; Sharma, N.; Albarrati, A.; Albratty, M.; Makeen, H.A.; Najmi, A.; Verma, R.; Bungau, S.G. Exploring the multifocal role of phytoconstituents as antidepressants. Prog. Neuropsychopharmacol. Biol. Psychiatry 2023, 123, 110693. [Google Scholar] [CrossRef]
  216. Bahramsoltani, R.; Farzaei, M.H.; Farahari, M.S.; Rahini, R. Phytochemical constituents as future antidepressants: A comprehensive review. Rev. Neurosci. 2015, 26, 699–719. [Google Scholar] [CrossRef]
  217. Al-Karawi, D.; Al Moori, D.A.; Tayyar, Y. The role of curcumin administration in patients with major depressive disorder: Mini metal-analysis of clinical trials. Phytother. Res. 2016, 30, 175–183. [Google Scholar] [CrossRef]
  218. Daubert, E.A.; Condron, B.G. Serotonin: A regulator of neuronal morphology and circuity. Trends Neurosci. 2010, 33, 424–434. [Google Scholar] [CrossRef]
  219. Lee, G.; Bae, H. Therapeutic effects of phytochemicals and medicinal herbs on depression. BioMed Res. Int. 2017, 2017, 6596241. [Google Scholar] [CrossRef]
  220. Phootha, N.; Yongparnichkul, N.; Fang, Z.X.; Gan, R.Y.; Zhang, P.Z. Plants and phytochemicals potentials in tackling anxiety: A systematic review. Phytomedicine Plus 2022, 2, 100375. [Google Scholar] [CrossRef]
  221. Andrade, S.; Nunes, D.; Dabur, M.; Ramalho, M.J.; Pereira, M.C.; Loureiro, J.A. Therapeutic potential of natural compounds in neurodegenerative diseases: Insights from clinical Trials. Pharmaceutics 2023, 15, 212. [Google Scholar] [CrossRef]
  222. de Lima, E.P.; Laurindo, L.F.; Catharin, V.C.S.; Direito, R.; Tanaka, M.; Jasmin Santos German, I.; Lamas, C.B.; Guiguer, E.L.; Araújo, A.C.; Fiorini, A.M.R.; et al. Polyphenols, alkaloids, and terpenoids against neurodegeneration: Evaluating the neuroprotective effects of phytocompounds through a comprehensive review of the current evidence. Metabolites 2025, 15, 124. [Google Scholar] [CrossRef]
  223. Crews, W.; Harrison, D.; Griffin, M.; Addison, K.; Yount, A.; Giovenco, M.; Hazell, J. A double-blinded, placebo-controlled, randomized trial of the neuropsychologic efficacy of cranberry juice in a sample of cognitively intact older adults: Pilot study findings. J. Altern. Complement. Med. 2005, 11, 305–309. [Google Scholar]
  224. Flanagan, E.; Cameron, D.; Sobhan, R.; Wong, C.C.; Pontifex, M.G.; Tosi, N.; Mena, P.; Del Rio, D.; Sami, S.A.; Narbad, A.; et al. Chronic consumption of cranberries (Vaccinium macrocarpon) for 12 weeks improves episodic memory and regional brain perfusion in healthy older adults: A randomised, placebo-controlled, parallel-groups feasibility study. Front. Nutr. 2022, 9, 849902. [Google Scholar] [CrossRef]
  225. Battaglia, S.; Avenanti, A.; Vécsei, L.; Tanaka, M. Neurodegeneration in cognitive impairment and mood disorders for experimental, clinical and translational neuropsychiatry. Biomedicines 2024, 12, 574. [Google Scholar] [CrossRef]
  226. Wang, Y.; Hernandez, G.M.D.; Mack, W.J.; Schneider, L.S.; Yin, F.; Brinton, R.D. Retrospective analysis of phytoSERM for management of menopause-associated vasomotor symptoms and cognitive decline: A pilot study on pharmacogenomic effects of mitochondrial haplogroup and APOE genotype on therapeutic efficacy. Menopause 2020, 27, 57–65. [Google Scholar] [CrossRef] [PubMed]
  227. Colombage, R.L.; Holden, S.; Lamport, D.; Barfoot, K.L. The effects of flavonoid supplementation on the mental health of postpartum parents. Front. Glob. Women’s Health 2024, 5, 1345353. [Google Scholar] [CrossRef]
  228. Kabra, A.; Garg, R.; Brimson, J.; Živković, J.; Almawash, S.; Ayaz, M.; Nawaz, A.; Hassan, S.S.U.; Bungau, S. Mechanistic insights into the role of plant polyphenols and their nano-formulations in the management of depression. Front. Pharmacol. 2022, 13, 1046599. [Google Scholar] [CrossRef]
  229. Shariatgorji, M.; Nilsson, A.; Goowin, R.J.A.; Källback, P.; Schintu, N.; Zhang, X.; Crossman, A.R.; Bezard, E.; Svenningsson, P.; Andren, P.E. Direct targeted quantitative molecular imaging of neurotransmitters in brain tissue sections. Neuron 2014, 84, 697–707. [Google Scholar] [CrossRef]
  230. Makkonen, I.; Riikonen, R.; Kokki, H.; Airaksinen, M.M.; Kuikka, J.T. Serotonin and dopamine transporter binding in children with autism determined by SPECT. Dev. Med. Child. Neurol. 2008, 50, 593–597. [Google Scholar] [CrossRef]
  231. Fowler, J.S.; Logan, J.; Shumay, E.; Alia-Klein, N.; Wang, G.J.; Volkow, N.D. Monoamine oxidase: Radiotracer chemistry and human studies. J. Label. Compd. Radiopharm. 2015, 58, 51–64. [Google Scholar] [CrossRef]
  232. Sacher, J.; Houle, S.; Parkes, J.; Rusjan, P.; Sagrati, S.; Wilson, A.A.; Meter, J.H. Monoamine oxidase A inhibitor occupancy during treatment of major depressive episodes with moclobemide or St. John’s wort: An [11C]-harmine PET study. J. Psychiatry Neurosci. 2011, 36, 375–382. [Google Scholar] [CrossRef] [PubMed]
  233. Manach, C.; Wikkiamson, G.; Morand, C.; Scalbert, A.; Remesy, C. Bioavailability and bioefficacy of polyphenols in humans. I. Review of 97 bioavailability studies. Am. J. Clin. Nutr. 2005, 81 (Suppl. S1), 230S–242S. [Google Scholar] [CrossRef]
  234. Campisi, A.; Sposito, G.; Pellitteri, R.; Santonocito, D.; Bisicchia, J.; Raciti, G.; Russo, C.; Nardiello, P.; Pignatello, R.; Casamenti, F.; et al. Effect of unloaded and curcumin-loaded solid lipid nanoparticles on tissue transglutaminase isoforms expression levels in an experimental model of Alzheimer’s disease. Antioxidants 2022, 11, 1863. [Google Scholar] [CrossRef] [PubMed]
  235. Rassu, G.; Porcu, E.P.; Fancello, S.; Obinu, A.; Senes, N.; Galleri, G.; Migheli, R.; Gavini, E.; Giunchedi, P. Intranasal delivery of genistein-loaded nanoparticles as a potential preventive system against neurodegenerative disorders. Pharmaceutics 2019, 11, 8. [Google Scholar] [CrossRef]
  236. Bandiwadekar, A.; Jose, J.; Gopan, G.; Augustin, V.; Ashtekar, H.; Khot, K.B. Transdermal delivery of resveratrol loaded solid lipid nanoparticle as a microneedle patch: A novel approach for the treatment of Parkinson’s disease. Drug Deliv. Transl. Res. 2025, 15, 1043–1073. [Google Scholar] [CrossRef]
  237. Kisku, A.; Nishad, A.; Agrawal, S.; Paliwal, R.; Datusalia, A.K.; Gupta, G.; Singh, S.K.; Dua, K.; Sulakhiya, K. Recent developments in intranasal drug delivery of nanomedicines for the treatment of neuropsychiatric disorders. Front. Med. 2024, 11, 1463976. [Google Scholar] [CrossRef]
  238. Jithavech, P.; Suwattananuruk, P.; Hasriadi, M.C.; Thitikornpong, W.; Towiwat, P.; Vajragupta, O.; Rojsitthisak, P. Physicochemical investigation of a novel curcumin diethyl γ-aminobutyrate, a carbamate ester prodrug of curcumin with enhanced anti-neuroinflammatory activity. PLoS ONE 2022, 17, e0265689. [Google Scholar] [CrossRef]
  239. Ahmad, S.; Hafeez, A. Formulation and development of curcumin-piperine-loaded S-SNEDDS for the treatment of Alzheimer’s disease. Mol. Neurobiol. 2023, 60, 1067–1082. [Google Scholar] [CrossRef]
  240. Sree, N. Recent advances in drug delivery systems for targeting the blood-brain barrier: Review article. J. Pharma Insights Res. 2024, 2, 34–44. [Google Scholar] [CrossRef]
  241. Kitagawa, S. Inhibitory effects of polyphenols on p-glycoprotein-mediated transport. Biol. Pharm. Bull. 2006, 29, 1–6. [Google Scholar] [CrossRef] [PubMed]
  242. Campos-Bedolla, P.; Walter, F.R.; Vaszelka, S.; Deli, M.A. Role of the blood-brain barrier in the nutrition of the central nervous system. Arch. Med. Res. 2014, 45, 610–638. [Google Scholar] [CrossRef] [PubMed]
  243. Scheepens, A.; Tan, T.; Paxton, J.W. Improving the oral bioavailability of beneficial polyphenols through designed synergies. Genes. Nutr. 2010, 5, 75–87. [Google Scholar] [CrossRef]
  244. Chen, J.; Chen, J.; Yu, P.; Yang, C.; Xia, C.; Deng, J.; Yu, M.; Xiang, Z.; Gan, L.; Zhu, B.; et al. A Novel Quercetin Encapsulated Glucose Modified Liposome and Its Brain-Target Antioxidative Neuroprotection Effects. Molecules 2024, 29, 607. [Google Scholar] [CrossRef] [PubMed]
  245. Fernandes, M.; Lopes, I.; Magalhães, L.; Sárria, M.P.; Machado, R.; Sousa, J.C.; Botelho, C.; Teixeira, J.; Gomes, A.C. Novel concept of exosome-like liposomes for the treatment of Alzheimer’s disease. J. Control. Release 2021, 336, 130–143. [Google Scholar] [CrossRef]
  246. Zhang, Q.L.; Fu, B.M.; Zhang, Z.J. Borneol, a novel agent that improves central nervous system drug delivery by enhancing blood–brain barrier permeability. Drug Deliv. 2017, 24, 1037–1044. [Google Scholar] [CrossRef]
Figure 1. Biosynthesis and metabolism of 5-HT. TPH hydroxylates L-tryptophan to 5-hydroxytryptophan (5-HTP), which AADC decarboxylates to 5-HT. MAO-A oxidizes 5-HT to its aldehyde, which ALDH oxidizes to 5-hydroxyindolacetic acid (5-HIAA). TPH requires BH4 and O2 for hydroxylation (H2O) and quinonoid dihydrobiopterin (q-BH2).
Figure 1. Biosynthesis and metabolism of 5-HT. TPH hydroxylates L-tryptophan to 5-hydroxytryptophan (5-HTP), which AADC decarboxylates to 5-HT. MAO-A oxidizes 5-HT to its aldehyde, which ALDH oxidizes to 5-hydroxyindolacetic acid (5-HIAA). TPH requires BH4 and O2 for hydroxylation (H2O) and quinonoid dihydrobiopterin (q-BH2).
Ijms 26 02916 g001
Figure 2. Biosynthesis pathway of catecholamine neurotransmitters in the brain. TH hydroxylates L-tyrosine to L-DOPA, which AADC decarboxylates to DA. DBH hydroxylates DA into NE, which is methylated to E by PNMT. DA are oxidized by MAO-A and MAO-B to the corresponding aldehydes and further oxidized by ALDH and O-methylates by COMT.
Figure 2. Biosynthesis pathway of catecholamine neurotransmitters in the brain. TH hydroxylates L-tyrosine to L-DOPA, which AADC decarboxylates to DA. DBH hydroxylates DA into NE, which is methylated to E by PNMT. DA are oxidized by MAO-A and MAO-B to the corresponding aldehydes and further oxidized by ALDH and O-methylates by COMT.
Ijms 26 02916 g002
Figure 3. Classification and chemical structures of major phytochemicals discussed in the text.
Figure 3. Classification and chemical structures of major phytochemicals discussed in the text.
Ijms 26 02916 g003
Figure 4. Structures required for antioxidant actions of flavonoids. (I); for radical scavenging (II); for metal chelating. The catechol group (3′, 4′-hydroxyl group) in the ring B, 2,3-double bond in conjugation with a 4-keto function in ring C and 3- and 5-hydroxyl group in rings C and A are essential structures for antioxidant function.
Figure 4. Structures required for antioxidant actions of flavonoids. (I); for radical scavenging (II); for metal chelating. The catechol group (3′, 4′-hydroxyl group) in the ring B, 2,3-double bond in conjugation with a 4-keto function in ring C and 3- and 5-hydroxyl group in rings C and A are essential structures for antioxidant function.
Ijms 26 02916 g004
Figure 5. Estrogen binds to ERβ, forming a complex that activates ERE at the TPH2 or TH promoter, or binds to Sp1, AP-1, CCAAT/enhancer binding protein b (C/EBPβ), or NF-κB, then to their binding sites, promotes the expression. Estrogen binds to GPER1, activates protein kinases cascade and transcription factors. Estrogen binds ERE and downregulates MAO-A transcription. Sex-determining region also regulates MAO-A transcription. Stress increased glucocorticoid and activated MAO-A expression. Phytoestrogens with high affinity to ERβ are shown.
Figure 5. Estrogen binds to ERβ, forming a complex that activates ERE at the TPH2 or TH promoter, or binds to Sp1, AP-1, CCAAT/enhancer binding protein b (C/EBPβ), or NF-κB, then to their binding sites, promotes the expression. Estrogen binds to GPER1, activates protein kinases cascade and transcription factors. Estrogen binds ERE and downregulates MAO-A transcription. Sex-determining region also regulates MAO-A transcription. Stress increased glucocorticoid and activated MAO-A expression. Phytoestrogens with high affinity to ERβ are shown.
Ijms 26 02916 g005
Figure 6. Mechanism of diet components and phytochemicals in the regulation of epigenetic modification. Folate, vitamin B12, and genistein increase SAM availability. Curcumin, apigenin, genistein, catechin, EGCG, and luteolin inhibit DNMT, HAT and HDAC. Fisetin, silibinin, and daidzein activate HDAC, leading to regulate gene expression. Some phytochemicals upregulate miRNA or downregulate lncRNA.
Figure 6. Mechanism of diet components and phytochemicals in the regulation of epigenetic modification. Folate, vitamin B12, and genistein increase SAM availability. Curcumin, apigenin, genistein, catechin, EGCG, and luteolin inhibit DNMT, HAT and HDAC. Fisetin, silibinin, and daidzein activate HDAC, leading to regulate gene expression. Some phytochemicals upregulate miRNA or downregulate lncRNA.
Ijms 26 02916 g006
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

Naoi, M.; Wu, Y.; Maruyama, W.; Shamoto-Nagai, M. Phytochemicals Modulate Biosynthesis and Function of Serotonin, Dopamine, and Norepinephrine for Treatment of Monoamine Neurotransmission-Related Psychiatric Diseases. Int. J. Mol. Sci. 2025, 26, 2916. https://doi.org/10.3390/ijms26072916

AMA Style

Naoi M, Wu Y, Maruyama W, Shamoto-Nagai M. Phytochemicals Modulate Biosynthesis and Function of Serotonin, Dopamine, and Norepinephrine for Treatment of Monoamine Neurotransmission-Related Psychiatric Diseases. International Journal of Molecular Sciences. 2025; 26(7):2916. https://doi.org/10.3390/ijms26072916

Chicago/Turabian Style

Naoi, Makoto, Yuqiu Wu, Wakako Maruyama, and Masayo Shamoto-Nagai. 2025. "Phytochemicals Modulate Biosynthesis and Function of Serotonin, Dopamine, and Norepinephrine for Treatment of Monoamine Neurotransmission-Related Psychiatric Diseases" International Journal of Molecular Sciences 26, no. 7: 2916. https://doi.org/10.3390/ijms26072916

APA Style

Naoi, M., Wu, Y., Maruyama, W., & Shamoto-Nagai, M. (2025). Phytochemicals Modulate Biosynthesis and Function of Serotonin, Dopamine, and Norepinephrine for Treatment of Monoamine Neurotransmission-Related Psychiatric Diseases. International Journal of Molecular Sciences, 26(7), 2916. https://doi.org/10.3390/ijms26072916

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