Next Article in Journal
Aloperine Inhibits ASFV via Regulating PRLR/JAK2 Signaling Pathway In Vitro
Previous Article in Journal
Monitoring Fruit Growth and Development in Apricot (Prunus armeniaca L.) through Gene Expression Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Biology and Biochemistry of Kynurenic Acid, a Potential Nutraceutical with Multiple Biological Effects

by
Luana de Fátima Alves
1,*,
J. Bernadette Moore
2,3 and
Douglas B. Kell
1,3,*
1
The Novo Nordisk Foundation Center for Biosustainability, Technical University of Denmark, Building 220, Søltofts Plads, 2800 Kongens Lyngby, Denmark
2
School of Food Science & Nutrition, University of Leeds, Leeds LS2 9JT, UK
3
Department of Biochemistry, Cell & Systems Biology, Institute of Systems, Molecular and Integrative Biology, University of Liverpool, Crown St., Liverpool L69 7ZB, UK
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(16), 9082; https://doi.org/10.3390/ijms25169082 (registering DOI)
Submission received: 19 July 2024 / Revised: 16 August 2024 / Accepted: 19 August 2024 / Published: 21 August 2024
(This article belongs to the Section Bioactives and Nutraceuticals)

Abstract

:
Kynurenic acid (KYNA) is an antioxidant degradation product of tryptophan that has been shown to have a variety of cytoprotective, neuroprotective and neuronal signalling properties. However, mammalian transporters and receptors display micromolar binding constants; these are consistent with its typically micromolar tissue concentrations but far above its serum/plasma concentration (normally tens of nanomolar), suggesting large gaps in our knowledge of its transport and mechanisms of action, in that the main influx transporters characterized to date are equilibrative, not concentrative. In addition, it is a substrate of a known anion efflux pump (ABCC4), whose in vivo activity is largely unknown. Exogeneous addition of L-tryptophan or L-kynurenine leads to the production of KYNA but also to that of many other co-metabolites (including some such as 3-hydroxy-L-kynurenine and quinolinic acid that may be toxic). With the exception of chestnut honey, KYNA exists at relatively low levels in natural foodstuffs. However, its bioavailability is reasonable, and as the terminal element of an irreversible reaction of most tryptophan degradation pathways, it might be added exogenously without disturbing upstream metabolism significantly. Many examples, which we review, show that it has valuable bioactivity. Given the above, we review its potential utility as a nutraceutical, finding it significantly worthy of further study and development.

1. Introduction

Many natural products, including normal human metabolites, are of interest as candidate nutraceuticals since their deficiency, while not necessarily causing overt disease, may lead to a less-than-optimal functioning of the organism of interest [1]. Accordingly, improved functioning, and the potential for an extended and healthy lifespan, might then be realized by the addition of the nutraceutical. Research interest in such nutraceuticals, which when added to and delivered in food matrices are referred to as ‘functional foods’, is consequently considerable (e.g., [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28]).
As part of a continuing survey of nutraceuticals, where we previously focused on ergothioneine (e.g., [18,29]), we determined that kynurenic acid (KYNA) might be of nutraceutical value. Here, we bring together some of the evidence that leads us to suppose that given its somewhat limited availability in normal diets, not least as the end product of a mammalian metabolic pathway, KYNA might indeed have a nutraceutical effect when provided exogenously.

2. Discovery, Structure and Some Biophysical Properties

Kynurenic acid (quinurenic acid, 4-oxo-1,4-dihydroquinoline-2-carboxylic acid, or 4-hydroxyquinoline-2-carboxylic acid) (https://pubchem.ncbi.nlm.nih.gov/substance/4854, accessed on 19 August 2024) was first identified in the urine of dogs by Justus von Liebig in 1853 [30]. It can adopt both keto and enol tautomers, as illustrated in Figure 1.
In aqueous solution at neutral pH, the keto form predominates [31,32,33], although a variety of crystal polymorphs are known [34]. It exhibits modest aqueous solubility (XlogP = 1.3) (pI~2.1 [35]), is very stable thermally [36,37], and can act as a photosensitizer [33,38,39] and metal chelator [40,41,42,43].

3. Biosynthesis and Phylogenetic Distribution

In humans, tryptophan is one of nine essential proteinogenic amino acids that must be obtained from the diet (i.e., we do not biosynthesize it [44]), while it (and its metabolites such as KYNA) may also be produced by the gut microbiome [45,46,47,48] (and transported in a manner reflected in the gut–brain axis [49,50,51,52,53,54,55,56,57,58,59,60]) or by some other minor pathways [61]. More than 95% of tryptophan is said to be degraded via the “L-kynurenine pathway” (KP) [62,63,64], consistent with mathematical models [65], and KYNA is one terminal part of the tryptophan degradation pathway, occurring via N-formylkynurenine and L-kynurenine (KYN) [66]. Note, however, that not all arms of this pathway exist in all cells: they are often segregated [67,68,69,70,71,72], implying the need for single-cell analyses [73,74,75,76,77,78,79]. While L-kynurenine is of interest as it is a precursor of NAD+ in eukaryotes (Figure 2) [80,81], our focus is on KYNA, made via the terminal stages of the pathway leading to KYNA from tryptophan (Figure 3), which goes via N-formyl kynurenine and L-kynurenine, and thus consists of three enzymes. Depending on the exact organism [82], these are tryptophan dioxygenase/indole dioxygenase (EC 1.13.11.11 and 1.13.11.52), kynurenine formamidase (E.C. 3.5.1.9) [83], and kynurenine oxoglutarate transaminase (E. C. 2.6.1.7). We note in passing that some members of the kynurenic acid pathway such as quinolinic acid [84,85] and 3-hydroxy-L-kynurenine [86] are considered neurotoxic (and KYNA can overcome this toxicity [87,88,89,90,91,92,93,94,95,96]) and so adding upstream elements, or elements that are more or less in equilibrium with them, is not necessarily a good idea. Additionally, KAT reactions in humans are normally considered to be of relatively minor significance due to the higher Km for its two substrates (in the millimolar range), when compared with the Km for L-kynurenine of the other two competitive reactions (catalyzed by kynurenine monooxygenase (KMO) and kynureninase A), that are in the micromolar range [97,98]. However (and see below), since KYNA is both seen as neuroprotective (e.g., [93,99,100]) and is essentially the terminal element and an irreversible step in this part of the kynurenine pathway, it is reasonable that provided it and its metabolites are beneficial or at least harmless, it can be added with impunity. Importantly, no diseases seem to be associated with the overexpression of KAT [101], the enzyme that is responsible for the synthesis of KYNA. The ability to add KYNA without affecting levels of molecules such as L-kynurenine directly is a core idea behind its potential use as a nutraceutical.

4. The Metabolic Pathway from Tryptophan to KYNA

Although a proper assessment of the flux to KYNA using metabolic control analysis (see [103,104,105,106,107]) seems not to have been performed, ODE-based mathematical modelling of the overall pathway has been [108], the interactions with vitamin B6 (pyridoxamine/pyridoxal) being of particular interest from a nutritional point of view as this is a cofactor for the KAT reaction. Readers are referred to the article [108] for a summary of the enzyme kinetic parameters of this pathway in mammals. Notably, both the Km and kcat values of the KAT enzyme(s) responsible for the production of KYNA from L-kynurenine are both rather high, indicating a tendency for linear increases in KYNA concentrations as that of L-kynurenine is raised [109].
Another ODE model for tryptophan degradation, notably in the liver, is given by Stavrum et al. [65], available at https://www.ebi.ac.uk/biomodels/MODEL1310160000 (accessed on 19 August 2024) and https://www.ebi.ac.uk/biomodels/BIOMD0000000602 (accessed on 19 August 2024), though the levels of KYNA are not reported in the paper. Importantly, the Copasi [110,111,112] files in SBML [113] version of the model available at https://www.ebi.ac.uk/biomodels/MODEL1310160000#Files (accessed on 19 August 2024) also indicate (i) that the KAT(1-4) reactions are seen as irreversible, and (ii) the flux-control coefficient (0.95) of the KAT steps on the flux towards KYNA is really dominating here. Of course, the values used may be varied, and a summary of kinetic parameters from different systems is given in Supplementary Information (Supplementary Table S1). Tissue distributions are tabulated in Supplementary Table S2.
Note too that a structural metabolic network model is also available as part of Recon2 [114] (https://www.ebi.ac.uk/biomodels/MODEL1311110001, accessed on 19 August 2024) and see [115]). Recon2 is a consensus network reconstruction based on the strategy used in [116] to produce one in baker’s yeast. Because just three enzymes catalyze the flux from the main dietary source of KYNA (viz L-tryptophan), we consider it worthwhile to review their properties in broad outline.

4.1. Tryptophan Dioxygenase/Indole Dioxygenase (EC 1.13.11.11 and 1.13.11.52)

Tryptophan dioxygenases (TDO) and indole dioxygenases (IDO) are heme-containing enzymes involved in the initial (and what is considered to be the most rate-limiting) step of the KP, catalyzing the oxidative cleavage of the indole ring of L-Trp to produce NFK (Figure 3) [117]. While IDO are widely distributed in the metazoan, in many bacterial species and fungi, and more recently have been identified in choanoflagellates and some ciliate species [118,119,120,121,122,123], TDO are also present in metazoan, bacteria and choanoflagellates, but have not been identified in fungi [124]. On the other hand, multiple IDO isoforms have been identified in fungal species [119,120]. Among them, IDOα isoforms usually show the lower Km values while IDOβ isoforms show higher Km but higher reaction velocities, resulting in higher catalytic efficiencies and suggesting that IDOβ could functionally substitute IDOα in fungal L-Trp metabolism for NAD+ production [120]. A third fungal IDO isoform, IDOγ, generally shows very low enzymatic activity for L-Trp (with catalytic efficiencies ranging around 1/100 of those determined for IDOα and IDOβ); however, IDOγ is very well conserved in fungi, suggesting that it might play an important role in those organisms, beyond NAD+ production [120].
Similarly, and apart from TDO, two distinct IDO genes have been identified in vertebrates. In humans, IDO1 and IDO2 are encoded by genes located in tandem on chromosome 8, which suggests gene duplication during evolution [125,126]. Although IDO and TDO catalyze the same reaction, there are fundamental differences between their structures, substrate specificity, tissue distribution and, consequently, function (see Supplementary Information).
In humans, IDO1 (hIDO1—UniProt P14902) has 403 amino acids, is monomeric, and with the exception of the liver, is widely distributed and constitutively expressed in multiple different tissues in healthy conditions, including lungs, small intestine, lymphatic system organs, female reproductive organs and the placenta [127,128]. Additionally, hIDO1 is constitutively expressed in several tumor cells, what is considered to make it a potential candidate for targeted anti-cancer therapy [129,130].
Under normal physiological conditions, IDO1 plays a fundamental role in immune regulation, acting as a checkpoint for the modulation of immune response mediated by antigen-presenting cells and exerting an immunosuppressive function, mediating maternal-fetal tolerance and protecting the fetus from maternal immune rejection [131]. In most cell types, however, hIDO1 is not only expressed constitutively under normal physiological conditions, but the enzyme is strongly induced in response to inflammation and infection stimuli, with IFN-γ being the main inducer [132,133], making IFN-γ-mediated hIDO1 induction and local L-Trp depletion an important factor able to inhibit pathogen growth [134].
In terms of enzyme activity, hIDO1 has the highest affinity for L-Trp (Km~20 μM) when compared with hIDO2 and hTDO, and it is also able to catalyze the oxygenation of D-Trp (Kcat~2.97 s−1), although the Km value for L-Trp is >100-fold higher than for D-Trp, suggesting a much weaker binding for D-Trp. Additionally, hIDO1 has a broad substrate specificity, catalyzing the oxygenation of a variety of indoleamines such as 5-hydroxytryptophan, 1-methyltryptophan, 5-methyltryptophan and 5-fluorotryptophan [135] and even serotonin in other organisms [136]. The activation of IDO1 activity requires a one-electron reduction of the heme—from the ferric (FeIII) state to the ferrous (FeII) state—facilitating the binding of O2 and L-Trp to the ternary complex [135]. The superoxide anion radical (O2●−) is the reducing cofactor and co-substrate for the purified IDO1 enzyme [137].
The human homolog of hIDO1, hIDO2 (UniProt Q6ZQW0), is also a monomeric protein with 407 amino acids that displays enzymatic activity towards L-Trp, although with a much higher Km, around 6.8 mM, when compared with hIDO1 and hTDO [138,139]. This value is more than 100-fold higher than typical physiological L-Trp levels [140,141], making it questionable if IDO2 plays a direct role in L-Trp metabolism. Yuasa and Ball showed that hIDO2 expression in Saccharomyces cerevisiae strains auxotrophic for nicotinic acid was not able to rescue the auxotrophic phenotype in the yeast, while expression of hIDO1 was, suggesting that the lower activity of hIDO2 might not be enough for NAD+ synthesis in yeast [121].
hIDO2 is not well characterized and little is known about its tissue distribution—at least at the protein level—and function. That is mainly due to (i) the complexity of hIDO2 transcription and, (ii) the lack of an accurately validated antibody. The hIDO2 gene generates five alternative transcripts, of which only one encodes the full-length protein [126]. Additionally, the gene contains two functional polymorphisms in the coding sequence: the first one, a nonsynonymous substitution (R248W) reduces hIDO2 catalytic activity by ~90% and the second one, a premature stop codon (Y359X), completely abolishes it. These polymorphisms have high prevalence in some populations—up to 50% [126].
Full length hIDO2 mRNA was detected in placenta and brain by RT-PCR, while primers specific for the hIDO2 exon 10 (common to all hIDO2 transcript forms) detected hIDO2 mRNAs in the liver, intestine, thymus, lung, spleen and kidney [126]. At the protein level, a few studies have identified hIDO2 in lungs, dendritic cells and in the interface between the placenta and the fetus [142,143,144].
By contrast to hIDO2, mouse IDO2 (mIDO2) is better studied. The constitutive expression of mIDO2 was detected in many organs at the protein and mRNA level (see Supplementary Table S1) and additionally, mIDO2 mRNA was upregulated in dendritic cells and mesenchymal stem cells treated with IFN-γ [132]. A recent study showed that IDO2 mediates autoreactive B-cell responses in mice, contributing to an exaggerated inflammatory response and the severity of joint inflammation in a model of autoimmune arthritis [145].
Unlike IDO1 and IDO2, hTDO (UniProt P48775) is a tetrameric enzyme [146] that is primarily confined in the liver and brain, where it seems to remain unresponsive to immunological stimuli; therefore, functioning as the main regulator of systemic tryptophan levels under physiological conditions [147]. As the main ‘housekeeping’ (see [148]) enzyme responsible for metabolizing the dietary tryptophan, hTDO is upregulated when the blood concentration of tryptophan rises above ‘physiological’ levels [149]. hTDO is well studied as a potential drug target, as its mRNA expression appears to be upregulated in many tumor types [150]. As it is mainly expressed in the liver and brain, human hepatocarcinoma usually present increased hTDO expression and many studies have suggested the involvement of hTDO in CNS diseases such as Alzheimer’s, Parkinson’s and Huntington’s disease [151,152,153].
Another difference of hTDO when compared to hIDO is that hTDO has a high substrate specificity, L-Trp being the only relevant physiological substrate, although there is evidence for oxidation of D-Trp, but with a very low activity, (low kcat at the high concentration of D-Trp tested) [154]. Additionally, the binding affinity of L-Trp to the ferric and ferrous forms of hTDO is very similar, suggesting that hTDO does not specifically favor substrate binding to the ferrous enzyme, as observed for hIDO [154].

4.2. Kynurenine Formamidase (E.C. 3.5.1.9) [83]

Human kynurenine formamidase (KF) (UniProt Q63HM1) or arylformamidase (Afmid) has 303 amino acids and catalyzes the second step of the KP from L-trp to KYNA by converting NFK to KYN (Figure 3). Its Alphafold-calculated structure is available, although very little is known about the human KF (hKF).
Pabarcus and Casida predicted [83] and later identified [155] the catalytic triad of the mouse KF (mKF). They showed that point and combined mutations in the Ser162, Asp247, and His279 (S164, D147 and H279 in the human protein—our unpublished alignment using the tools provided in [156]) completely abolished the conversion of NFK to KYN [155].
In mice, mKF is primarily expressed in the liver and to a less extent in the kidney [157]. Afmid knockout mice showed elevated plasma concentrations of NFK, KYN and KYNA (as well as kidney failure), consistent with low levels of mKF expression [158]. In S. cerevisiae, a KF knockout strain showed an accumulation of NFK and a slow growth phenotype in absence of exogenous nicotinate [159]. A range of compounds, including organophosphate and methylcarbamate insecticides are potent inhibitors of KF [160,161], and in vivo treatment of mouse with organophosphorus acid triester diazinon resulted in the accumulation of NFK and reduced KYN biosynthesis in the liver and a 5-fold increased plasma KYN and 5- to 15-fold increased concentrations of KYNA and xanthurenic acid in urine [162]. This suggested a strategy for the development of safer insecticides of this type [163].

4.3. Kynurenine Oxoglutarate Transaminase (E. C. 2.6.1.7)/Kynurenine Aminotransferase (KATs)

Overall, the simple pathway structure alone, plus other observable properties such as the modelling above and correlations between KAT levels and KYNA concentrations, leads one to suppose that this reaction, as catalyzed by various KATs (kynurenine aminotransferase), is especially important to the synthesis of KYNA. The reaction (BRENDA, https://www.brenda-enzymes.org/enzyme.php?ecno=2.6.1.7, accessed on 19 August 2024) has been shown to be effectively irreversible [164,165] in the direction of KYNA synthesis, and is treated as such in the ODE models [65]. The transamination of kynurenine by KATs yields an unstable keto acid product, 4-(2-aminophenyl)-2,4-dioxobutanoate, which is spontaneously and rapidly cyclized to KYNA [82,166] (Figure 3). This, importantly, is what makes this step functionally irreversible. KAT orthologous are found in all kingdoms [166], and a cross-species comparison of KAT structures shows a high conservation of the monomer architecture, consisting of an N-terminus arm, a small and a large domain [66,167,168].
Depending on the microbe, the function of KYNA is unclear, as inhibiting its significant production by inhibiting the relevant (Aro8/9) KAT enzymes in yeast, for example, has no significant effects [169]; arguably this enzyme activity may help detoxify excess tryptophan [169] (and excess amino acids can certainly be toxic to microbes [170]). By contrast, Genestet et al. observed that Pseudomonas aeruginosa clinical isolates present a high transcription level of the kynaA gene (the first gene involved in the kynurenine pathway, converting L-Trp to L-kynurenine) and produce high amounts of kynurenine when in contact with human neutrophils, leading to increased bacterial survival. By testing kynurenine-overproducing (ΔkynU—gene involved in the conversion of L-kynurenine into L-anthranilate) and kynurenine-deficient (ΔkynA) P. aeruginosa, they determined that L-kynurenine inhibits ROS production by neutrophils, but when testing the specific mechanisms, they failed in showing that L-kynurenine had a direct effect on NADPH oxidase (the main ROS producer in neutrophils) or was a potent scavenger of superoxide anions in a superoxide-producing nonenzymatic PMS–NADH system. In contrast, KYNA appeared as the best scavenger of O2 among the molecules tested [171]. In this case, an increase in L-kynurenine production by P. aeruginosa, especially when the pathway branch leading to anthranilate production is suppressed, might lead to an increase in the production of KYNA [172], which appears to be the best scavenger, although more studies are necessary to define its role over L-kynurenine in P. aeruginosa survival. Furthermore, KATs can play an essential role in bacterial survival and amino acid synthesis/nitrogen assimilation due to their higher affinity and efficiency for other natural substrates over L-kynurenine [168,173].
There are four human KATs, summarized by Rossi [66] and structures are available, e.g., for KAT1 [174,175], KAT II [176,177,178,179], KAT III (as a homology model [180]) and KAT IV [181]. There are certain differences in substrate specificity, but all are homodimeric pyridoxal phosphate-dependent enzymes [165]. Because of their importance to endogenous KYNA synthesis, we consider each in turn.

4.4. KAT1

The main isoform of KAT1 (Uniprot Q16773) has 422 amino acids, a broad substrate specificity as an aminotransferase and also catalyzes β-lyase reactions using several cysteine S-conjugates as substrates. It exhibits a preference for glutamine as amino donor (see Supplementary Table S2) and was demonstrated to have aminotransferase activity towards 5-S-L-cysteinyldopamine, the cysteine S-conjugate of dopamine; this is significant as 5-S-L-cysteinyldopamine is neurotoxic and markedly increased in the substantia nigra of patients who died of Parkinson’s disease [182,183]. As judged by the protein atlas [184], KAT I is widely distributed in human tissue, including brain [185].
In vitro KAT-1 can use many α-keto acids as amino group acceptors and although it has detectable activity on oxaloacetate and pyruvate, the specific activity on these two α-keto acids is very slow, making unlikely that they are physiological substrates for human KAT I [186]. Additionally, hKAT1 conversion of L-kynurenine to KYNA (using α-ketobutyrate as α-ketoacid) is strongly inhibited by tryptophan, phenylalanine, glutamine and cysteine (2 mM of each amino acid inhibits over 50% of aminotransferase activity) and by indo-3-pyruvate (0.2 mM inhibits 50%) [186]. In vivo preference of L-glutamine over L-kynurenine in brain is discussed by Cooper et al. based on L-glutamine and L-kynurenine availability/concentrations and their Km and kcat for KAT I, suggesting that the capacity of KAT I to utilize L-glutamine is many orders of magnitude higher than the capacity of the enzyme to utilize L-kynurenine [187]. This makes it very unlikely that KAT-1 has a major role in KYNA production in the brain. Indeed, many studies have shown that KAT II is the main enzyme responsible for KYNA production in the brain; however, KAT II knockout in mice led to a reduction of 71% in the KYNA levels in the brain [188]. We note that Kapoor et al. [189] showed that the enzymatic activity of KAT I in the brain of patients with schizophrenia was altered, and suggested that the enzyme might play an important role in KYNA synthesis in the brain; however, those two facts do not follow, and the observation does not seem to have been followed up.

4.5. KAT II

KATII (Uniprot Q8N5Z0), 425 residues, has a very broad substrate specificity, albeit with a preference for glutamate L-kynurenine, a Km for kynureine of 4.7 mM and a kcat of 585 ·min−1 (ca 10 ·s−1) while a catalytic efficiency of 196.2 mM−1·min−1 for aminoadipate is reported [177]. Its mechanism of action is known in detail [165], its molecular dynamics simulations are given by [190], and its pharmaceutical inhibitors are summarized, e.g., by [178,191,192,193,194,195]. KAT II is considered the major source of KYNA in mammals, and is widely distributed, not least in the liver. The interest in developing pharmacological inhibitors comes from the somewhat variable data (Table 4, below) indicating that KYNA levels can sometimes be raised in various kinds of bipolar disorder [196]. From the perspective of this review, however, we think it more likely that any beneficial effects of such inhibition may be mediated via other parts of the tryptophan degradation pathway that are likely to change simultaneously (see Figure 3, and, e.g., [197,198]), and indeed we are not aware of any marketed drug for such disorders based on selective inhibition of KAT II enzymes.
Unlike KAT I, the conversion of L-kynurenine to KYNA by KAT II is not significantly inhibited by other amino acids [177]. Taken together, preference for aminoadipate and L-kynurenine and the lack of inhibition by other amino acids might explain the importance of KAT II in preventing neurotoxicity. KAT II is responsible for as much as 75% of KYNA synthesis in most brain areas [61,199] and its downregulation has been related to numerous brain diseases involving KYNA depletion [200]. KATII might also be involved in the regulation of brain levels of aminoadipate [201], a toxic metabolite for astrocytes in vitro and in vivo and part of lysine metabolism in the liver [201,202,203].
Regarding sequence and structure, KAT II has a predicted 22 N-terminus mitochondrial signal sequence, targeting the enzyme to the inner membrane of mitochondria [204]. KAT II enzymes do not belong to any of the previously existing fold type I aminotransferase groups from the α-family of PLP-dependent enzymes, but to a new subgroup called Iε. This occurs because hKAT II have a highly flexible N-terminal fraction, residues 15–33, that is able to move far from and closer to the active site upon substrate binding, thus accommodating different substrate sizes, which can explain its broad substrate specificity [177]. The swapping of the catalytic N-terminal region is unique in this subgroup of aminotransferases [109]. The conformation adopted by the N-terminal region in human KAT II is similar to the one observed in the N-terminus of members of PLP-dependent lyases [66], and in fact, hKAT II shows β-lyase activity towards various cysteine S-conjugates and β-chloro-D,L-alanine [205].
Both hKAT I and hKAT II are by far the most well studied human KATs. Both enzymes have been largely targeted in structure-based drug design aiming (we think unadvisedly) at the lowering of KYNA levels. Among the most potent KAT I inhibitors are the phenylhydrazone hexanoic acid derivatives and, for KAT II, the pyrazole compounds. The inhibitors showed effectiveness in reducing KYNA production followed by an improvement in alleviating cognitive dysfunction in animal models, but studies in humans are yet to be performed [206,207], and mechanisms are far from clear cut.

4.6. KAT III

KAT III (Uniprot Q6YP21), most closely related in sequence to KAT I (unpublished alignment search using the tools provided in [156]), has 454 amino acids and, like KAT I, a preference for glutamine as amino donor. Like KAT I, it is more or less ubiquitously distributed in humans and shares a similar intron-exon organization, except for the presence of exon 2, which encodes a 33-amino acid sequence that corresponds to the leader sequence for mitochondrial targeting. Exon 2 can be alternatively spliced in hKAT III, and thus the enzyme can be found in the cytoplasm or mitochondria [204,208]. The human KAT III has not been biochemically characterized, but the mouse KAT III (mKAT III), which shares 86.8% similarity and 83.7% identity with the hKAT III, has been fully characterized. Here, we discuss the biochemical/kinetic parameters of the mKAT III. Similarly to hKAT I, mKAT III has a preference for glutamine as amino donor. The catalytic efficiency for L-kynurenine is 92 min−1 mM−1 and the transamination of L-kynurenine to form KYNA is significantly inhibited by methionine, histidine and glutamine (~75%), leucine and cysteine (~50%) and phenylalanine (~25% inhibition) [209].

4.7. KAT IV

Better known as the mitochondrial glutamate-oxaloacetate aminotransferase 2 (GOT2), KAT IV (Uniprot P00505) has 430 amino acids and catalyzes the reaction of 2-oxoglutarate and L-aspartate to L-glutamate and oxaloacetate, playing an essential role in the malate-aspartate shuttle in mitochondria and in the synthesis of glutamate [210]. It is very widely distributed in human tissues, and an experimental [181] and Alphafold-calculated structure is available. KAT IV has been shown to play significant role on KYNA synthesis in human and murine brains [211]. Biochemical characterization of mouse mitochondrial KAT IV (mKAT IV) showed high transamination activity towards glutamate and aspartate as amino donors and lower but detectable activity towards phenylalanine, tyrosine, cysteine, tryptophan, 3-HK, methionine, kynurenine, and asparagine [109]. As amino group acceptors, mKAT IV showed similar Km values for phenylpyruvate and oxaloacetate, but higher catalytic efficiency towards phenyl-pyruvate (58 mM−1·min−1) [212]. Accordingly, the transamination of L-kynurenine to KYNA was significantly inhibited by glutamate and aspartate, and in lower amounts by cysteine, glutamine, phenylalanine, tryptophan and tyrosine [109]. Notably, KAT IV was the one that showed substantially increased activity following endurance exercise (as did KYNA) [213,214,215].
The overall conclusion here is that most tissues exhibit some basal KAT activity, consistent with the view that KYNA is a useful metabolite for mammals.

5. Transport of Kynurenic Acid and Related Metabolites

Mammalian kynurenic Acid Transporters SLC22A6 and SLC22A8

As is now well established, molecules such as KYNA require protein transporters to cross cell membranes [216,217,218,219,220,221,222,223,224,225,226,227,228]. Human transporters are classified into two superfamilies. The SLCs, for SoLute Carriers [229,230], are either equilibrative (effecting ‘facilitated diffusion’) or use ion electrochemical gradient to transport their substrates against ostensible concentration gradients (‘concentrative’). In addition, there are various ATP-binding cassette (ABC) transporter families, commonly encoding efflux transporters [231]. However, exceptions exist, and some are actually influx transporters [232,233,234]. The SLCs described to date as being involved in the transport of kynurenic acid are the related SLC22A6 and SLC22A8, which come from the ‘organic anion transport’ or OAT family [235,236] and were previously known as OAT1 and OAT3. They are polyspecific transporters with a very wide substrate range among anions [237], but are Na+-independent and not thought to be concentrative (unless balanced by an opposite starting concentration gradient of another substrate). Moreover, using the tissue data from [148,238], while their maximum expression profile levels are quite respectable among SLC22 family members (Figure 4), these gene products have an exceptionally high Gini coefficient (see [148,239]). This means that they are mainly expressed in a very small number of tissues (in this case the kidney, urinary bladder, and (for SLC22A8) brain (see e.g., https://www.proteinatlas.org/ENSG00000149452-SLC22A8/tissue, accessed on 19 August 2024), and not, for instance, the liver. This said, while it is stated that the blood–brain barrier itself is poorly permeable to KYNA [89], the original paper [240] on which the statement is based indicates that its rate of uptake is only ~10-fold lower than that of L-kynurenine (which, unlike KYNA, is a substrate of the LAT1/SLC7A5 transporter [241,242]). Correspondingly, we would argue that KYNA can in fact enter the brain if supplied exogenously, albeit the transporters (of which there may be many [243]) are as yet unknown. Seemingly, KYNA (and its 7-chloro derivative) can also be effluxed by MRCP4 (ABCC4) as well as by SLC22A6/8 [242,244], as each of these transporters is inhibited by probenecid, whose presence led to an 885-fold increase in the concentration of 7-Cl-KYNA in the prefrontal cortex of rats [242] (see also [240,245,246,247,248,249,250]). This may also help to account for the low apparent net uptake sometimes seen, as ABCC4 is a very efficient efflux pump [251], and efflux pumps in general necessarily tend to have more influence on steady-state levels of a drug than do influx transporters [252]. Indeed, they are a major cause of resistance to antibiotics (e.g., [232,253,254,255]) and to antitumor drugs (e.g., [256,257,258,259,260]). Therefore, the tissue distribution of KYNA is likely to depend strongly on the concentrations of potential ABCC4 inhibitors (some listed in Table 1) as well as the disposition of efflux transporters like ABCC4 and others of the ABCC family. Interestingly, among the most potent inhibitors of ABCC4 (Ki~1 μM [251]) is the flavonoid quercetin, another important nutraceutical [261,262]. ABCG2 (BCRP) is also a potential effluxer of KYNA [263,264]. Note too that kynurenic acid may also be bound to albumin [265,266], something that would be missed in standard extractive metabolomics studies. Overall, the system is extremely complex, benefitting strongly from the kind of ODE-based modelling that is known in this field as ‘physiologically based pharmacokinetic modelling’ [267,268], while the increasing availability of cell lines engineered to overexpress individual SLCs [269] should make answering this question of KYNA transporters much more accessible [270].
This probably means that, given also the much higher concentrations of KYNA in tissues versus plasma (see below), there are other kynurenic acid transporters waiting to be discovered. In this context, interestingly, a kynurenine monooxygenase inhibitor was transported via the riboflavin transporter SLC52A2 [85], and the structures of KYNA and riboflavin are given in Figure 5. There are also some weak indications [288] that kynurenates might be substrates of SLC1 [289,290] family members. Other obvious candidates are members of the SLC16 monocarboxylate transporter family [291,292,293]. KYNA transporters in other organisms are surprisingly poorly characterized. An especially striking finding [294] was that KYNA was accumulated 20-fold in cord blood relative to the maternal plasma, and its concentration was also 20-fold greater (in mice) in the fetal brain vs. maternal tissue [295] (see also [296] for external KYNA addition), strongly implying a role for both concentrative transporters and for KYNA itself in fetal development.
We note also that L-kynureine is more widely bioavailable, being a substrate of the ‘large amino acid transporter’ in C. elegans [297], a homologue of the human SLC7A5 [241,298,299,300,301,302] that transports many large, neutral amino acids, including tryptophan and L-kynurenine). This of course complicates analyses of the transport of KYNA if tryptophan or L-kynurenine and/or kynurenine amino-transferase(s) are also present or added (Figure 6). D-kynurenine can also be used and is metabolized via D-amino acid oxidase [303] or transamination [304].

6. Bioavailability

Given the relative paucity and activity of (known) transporters, it is possible that KYNA is not the most bioavailable of nutraceuticals, but that oral KYNA is definitely absorbed in the mammalian gut [36,305,306,307,308], and can cross a model of the blood–brain barrier with a respectable permeability of some 3.5·10−6 cm·s−1 [309], precisely that of the modal value for marketed drugs across Caco-2 cells [310]. The levels of natural absorption can also be improved further by linking it to a transporter substrate [311,312] or using special formulations [309,313,314,315,316,317,318] (we ignore analogues that do not yield KYNA itself, since our focus is the nutraceutical activity of the genuine natural product). Thus, there is no reason why exogenously supplied KYNA might not be bioavailable, and the many effects reviewed here surely indicate both that it is and that can confer host benefits.

7. Concentrations of KYNA in ‘Normal’ Serum and Plasma, and Other Body Fluids/Tissues

Concentrations in the plasma of the nutraceutical ergothioneine are typically 1–4 μM (~229–916 ng/mL) (e.g., [319,320]), and are ~10-fold higher in whole blood as ergothioneine is concentrated in erythrocytes [18]. In contrast, KYNA concentrations in plasma and serum are some 1–2 orders of magnitude lower (Table 2). Moreover, the KYNA plasma concentration is normally far lower than the ~5 μM Km values measured [321,322] for SLC22A6/8 (and indeed for many of its putative receptors—see [323] and below). The median levels are fairly consistently ~30–50 nM in plasma or serum across a very wide range of studies (Table 2). This relative constancy also implies a significant degree of regulation [324]. With a MW of 189 at pH 7, 50 nM equates to some 9.45 ng KYNA·mL−1 (one paper gives values of KYNA in children that are orders of magnitude different [325], and another [326] gives very unusually low values; these are not included in Table 2). A recent meta-analysis showing a tendency of serum/plasma KYNA to increase with age is given by [327], with similar data in [328,329], though tissue changes are rather variable [330].

7.1. Breast Milk

O’Rourke et al. measured various tryptophan metabolites including KYNA in human breast milk, finding values of ca. 12 ng/mL (52 nM) at term, rising ~3.5-fold for term babies, but less so (and not significantly) in pre-term babies. Milart et al. [359] noted that natural human breast milk contained KYNA at much higher levels (from ca. 21 nM to ~300 nM at 6 months of breastfeeding, a ca. 14-fold increase) as lactation kicked in, levels that were well above those in a variety of commercial infant formulas. Rat pups exposed postnatally to KYNA also demonstrated less obesity for the same increase in bone mineral density [359] (see also [360,361]).

7.2. Bile

Concentrations of KYNA in bile are well in excess of those in plasma/serum, with values of >800 nM [308,362] being reported in humans.

7.3. Intestine

Of course, depending on dietary sources of tryptophan and kynurenic acid, plus the variable ability of microbes in the gut to metabolize these molecules to KYNA, mean that KYNA levels could in some cases be quite high, and this has been reported (e.g., [334]).

7.4. Gut Microbiota and KYNA

The gut microbiota composition plays an important role in regulating KP metabolites, which subsequently influence host immune response. The interplay between these three is tightly controlled and complex; and the gut microbiota can influence health and disease through fine-tuning KP metabolites [50,52,363].
In the gastrointestinal (GI) tract, the function of AhR signalling as a critical regulator of gut immune function has been extensively reported [364,365,366,367,368,369]. Therefore, KYNA—as well as other members of the KP—may be an important mediator in this complex crosstalk, once it acts as a direct ligand of the above-mentioned receptor, activating it locally and systemically [51,370]. In fact, the absence of AhR causes an increase in endogenous KYNA levels in mice [93] and, in a recent study, gut microbiota-derived KYNA and other metabolites from tryptophan metabolism were shown to be the main AhR activators in the GI tract [371].
Another important role of ligand-activated AhR is the induction of IDO1 activity through activation of pro-inflammatory cytokines [372]. In this case, it is important to consider the influence of altered IDO activity on KP metabolite production [373].
In addition to AhR, transmembrane G protein-coupled receptors (GPCRs) also play an important role in GI tract homeostasis and intestinal immunity [374,375,376]. Among the GPCRs, GPR35 is predominantly expressed in the GI tract and, since KYNA is a known GPR35 ligand, several studies have suggested that KYNA acts as the link between gut-microbiota homeostasis and host immunological regulation. For example, Wang and collaborators demonstrated that GPR35 activation by KYNA is a necessary component to maintain gut homeostasis, regulating the progression and outcome of colitis in an ulcerative colitis-induced rat model [377]. Another study demonstrated that KYNA-mediated AhR and GPR35 regulation maintain intestinal integrity and homeostasis in a chemotherapeutics-induced intestinal damage model. Sensitivity differences of AhR and GRP35 to KYNA leads to a primary accumulation of KYNA through AhR-IDO1 positive feedback regulation. Accumulation of KYNA then is sensed by GPR35, which ameliorates intestinal injury and restores gut homeostasis [378]. Additionally, Miyamoto and collaborators have demonstrated [379] that the increased KYNA levels in the small intestine mediated by the microbiota modulates the recruitment and aggregation of GPR35-positive macrophages, ultimately triggering the onset of experimental autoimmune encephalomyelitis.
Tissues—Rodents
Concentration RangeCommentsReference
Review[347]
32 nM increases to → 135 mM after dosingGerbil brain [360]
1–16 mMRat ileum[361]
~40 nM in plasmaTrebled after dosing at 5 mg/kg[210]
Tissues—human
Concentration rangeCommentsReference
Review[305]
0.2–0.7 pmol/mgBrain; 3× increase in Down syndrome[362]
2–3 pmol/mgBrain[353]
Up to 1.58 pmol/mgBrain[363] and review [347]
1.6 mMColon[359]
10.2 ng/mLFetal membrane[364]
7.6 g/mLUmbilical Cord[364]
1 ng/mLPlacenta[364]
If we loosely assume a unit density (1 g·mL−1) for tissue, 1 pmol.mg−1 equates to 1 μM, considerably higher than serum/plasma levels, strongly implying that there is concentrative uptake driven via one or more (presumably H+- or Na+-coupled) transporters, whose identities—as with that of many SLCs [269]—remain unknown. Although it was assumed that it was the local rates of production that varied, this conclusion is also consistent with the analyses of maternal, fetal and cord blood in [380].

7.5. Urine

As reviewed by Turska et al. [308], urine concentrations of KYNA tend to be in the micromolar range, from ca. 4 μM [381] to more than 20 μM [382]. It is hard to know how much of this is due to simple synthesis from L-kynurenine in the kidney (for which there is no particular reason) and how much is due to concentrative efflux (noting again that SLC22A6/8 are considered to be exchangers [383,384] and not concentrative. Note that while SLC22A6 (OAT1) https://www.proteinatlas.org/ENSG00000197901-SLC22A6/tissue, accessed on 19 August 2024 and SLC22A8 (OAT3) https://www.proteinatlas.org/ENSG00000197901-SLC22A8/tissue, accessed on 19 August 2024 are highly expressed in the kidney, ABCC4 is not expressed in the kidney https://www.proteinatlas.org/ENSG00000125257-ABCC4/tissue, accessed on 19 August 2024.

7.6. Feces

These are somewhat infrequently measured, but in one rat study [385] levels were around 100 ng/g (~0.5 μM if 1 g ≡ 1 mL), rising to 40 times that in the presence of a kynurenine monooxygenase inhibitor.

8. Nutritional Sources

As a metabolite of an essential amino acid, KYNA is widely distributed, and plants can take it up from the soil [386]. The literature for natural products is focused on Ephedra spp. (e.g., [387,388]), which have a significant use in traditional Chinese medicine (MaHuang), but the contribution to this of KYNA is unknown and many Ephedra alkaloids can be toxic. Besides culinary herbs [55], where the richest sources are basil and thyme [307], or medicinal herbs that still might provide at most a few tens of μg [386,389], of those vegetables consumed in reasonable quantities, broccoli and potatoes seem to have the highest values (Table 3). As with many amino acids [390], the levels varied massively between different cultivars [391], warranting more detailed studies.
KYNA contents were also measured in tea and coffee, but did not exceed 8.7 and 0.63 μg/100 mL, respectively.
Honey is a notable source of KYNA. In particular, the product from sweet chestnut trees reportedly weighs in at ca. 100 mg KYNA/kg [36,307] or even more [392] (Table 3 and Figure 7). The source of these high levels is, in particular, the male flowers of the tree, most other parts of the edible chestnut having far lower levels [392]. Note, however, that most honeys are closer to 1 mg/kg or lower, so a 25 mg supplement (say) of KYNA would require a mighty dose of any but the most potent honey. An overall conclusion from Table 3 is that if KYNA is going to be given as a nutraceutical, even at low doses, its levels are likely to exceed those seen when its sole exogenous source is foodstuffs [308] (propolis, of an unstated origin, was also said to be a good source [36], although KYNA was not reported in a number of untargeted metabolomics studies [394,395,396,397], so this seems worth investigating further).
As with ergothioneine, where clear (even striking) benefits are seen from eating mushrooms that contain it, e.g., in preventing mild cognitive impairment [398], even though they may contain nutraceuticals beyond the one of focus, chestnut honey is clearly the equivalent for KYNA. Thus, while the precise contribution of KYNA itself is unknown (chestnut honey also contains many phenolic and other antioxidants [399]), the excellent review by Turska and colleagues [308] does provide a list of examples where this honey is thought to have provided health benefits. These include positive effects on glucose metabolism and neurodegeneration in obese mice [400], vs. high-fat diets in obese mice [401], acid-/alcohol-induced gastric ulceration in mice [402], on carbon tetrachloride-induced liver damage in rats [403], and in inhibiting breast cancer cell line proliferation in vitro [404]. It has been found protective against influenza in mouse macrophages and mice in vivo [405]. Along with curcumin, it also produced a substantial increase in the longevity of heat-stressed bees [406].

9. Pharmacokinetics

There have been few studies of the pharmacokinetics of exogenous KYNA [308], with Turska’s study in mice [407] being the stand out. Here, radiolabelled KYNA was provided intragastrally (calculated as ~5 nmol at 20 μM), and its appearance in blood, liver and spleen noted, indicating uptake into these organs. Kidney levels were not reported. Most of the KYNA was excreted in urine in under 24 h, while liver retained a significant level of radioactivity at that time. Note that liver in humans does not express the two known transporters (see above), nor significant amounts of the ABCC4 effluxer, implying the need for other, as yet unknown, transporters.

10. Further Metabolism and Excretion

In humans, KYNA is largely seen as a terminal step of tryptophan degradation [308], and as noted above is excreted in urine via the kidneys. As such, metabolic transformations are not considered a major feature of KYNA ingestions, though Takahashi et al. reported some conversion to quinaldic acid [305], the dehydroxylated variant of KYNA [408]. Various bacteria can of course metabolize it, e.g., certain pseudomonads can assimilate and metabolize it to glutamate, alanine and various organic acids [409], but they can also excrete it [172].

11. Oxidative Stress

Oxidative stress is extremely widespread in a whole host of chronic, inflammatory diseases [410,411], so much so that there were over 125 papers having the terms “oxidative stress” and “review” in their titles alone at Web of Knowledge just for 2023. Since we have reviewed elements of it three times recently [410,412,413], we do not repeat this further here, save to mention that the chief cause is the production of ‘reactive oxygen species’ (ROS) such as peroxide, superoxide, and—as catalysed by free iron [324,410,414,415]—the especially nasty hydroxyl radical OH. Any small antioxidant molecules that can react with such ROS (also known as ROS scavengers) are thus likely to ameliorate oxidative stress, and KYNA certainly has this property [416,417,418,419,420] (and see below). Below we discuss other mechanisms that may account for the ability of low concentrations of KYNA to help deal with oxidative stress.

12. Diseases in Which KYNA Levels Are Significantly Altered

Table 4 summarizes some of the diseases or syndromes in which normal levels of KYNA are raised (occasionally) or (more frequently [421]) lowered. In the former cases, there is some evidence that this is actually the host’s homoeostatic attempt to combat the causes of the disease. Note, of course, that in many cases it is probable that it is an increase in tryptophan and its degradation pathway metabolites more generally that are changed, so increases in levels upstream may themselves correlate with KYNA levels yet themselves be responsible for physiological or biochemical effects [422]. The importance of the kidney as the main means of excretion (via urine) is highlighted by the very high levels of plasma KYNA reached in various kidney diseases.

13. KYNA and COVID-19

The arrival of the COVID-19 epidemic caused by the SARS-CoV-2 virus has had profound effects on the world, as well as on scientific approaches to the understanding of both acute diseases and their post-acute or chronic sequelae (‘Long COVID’) [486,487]. A number of studies have highlighted changes in tryptophan metabolism and in KYNA production in particular as a response to the virus. Thus, Thomas et al. [488] found trp metabolism the most significant pathway changed statistically, while Roberts et al. [331,489] found L-kynurenine and KYNA among the metabolites most raised in terms of predicting both the severity of the disease and poor outcome (implying activation upstream of L-kynurenine as well as of KAT enzymes or transporters). Cihan et al. [490] and Kucukkarapinar et al. [491] reported similar data, while Sindelar and colleagues found KYNA to be the most predictive of disease severity [492]. Cai et al. suggested additional gender differences [493] (though we could not confirm that [331,489]), while L-kynurenine was noted by Almulla and colleagues but not KYNA [494]. Holmes and colleagues [495] inferred elements of the L-kynurenine pathway but reported only ratios. L-kynurenine was also noted by other authors, such as [496,497,498,499,500,501,502] (in these latter cases, KYNA was seemingly not measured, pointing up the utility of untargeted discovery methods for unravelling the biology more fully, both for metabolomics [503] and more generally [504]).
What is a priori unknown, however, for this or other diseases, is the extent to which the upstream metabolites such as L-kynurenine and quinolinic acid, considered to be less beneficial, counteract any possible benefits of KYNA, whether its production represents attempts by the body to use it as a protective agent, or whether it is simply a ‘by-product’ of L-kynurenine due to the presence of KAT activity; this needs testing with KYNA or KATs as an independent variable.

14. Protection against Various Diseases

Antioxidants such as ergothioneine are seen as excellent cytoprotectants against multiple stresses [505,506,507,508]. In a similar vein, KYNA has also been demonstrated to be a neuroprotectant and cytoprotectant against a variety of acute challenges. As before with ergothioneine [18], we divide these studies into central and peripheral studies, before looking at reported receptors for KYNA.

14.1. Neuroprotection

Many studies, some reviewed recently by [308,461], have indicated the ability of KYNA to serve as a neuroprotective agent, and some of the relevant papers are set out in Table 5.

14.2. Peripheral Protection

In Table 6, we summarize some of the examples in which KYNA has proved to be neuroprotective against stresses in non-CNS tissues. Ischemia-reperfusion injury occurs when tissues subjected to hypoxia are reoxygenated, leading to the rapid formation of ROSs; although this is well known in acute circumstances, it is becoming increasingly recognized that this can also occur chronically (e.g., [413]), especially in diseases such as long COVID where fibrinaloid microclots [539,540,541] can induce hypoxia [539] and related sequelae such as reperfusion injury [413] and postural orthostatic tachycardia syndrome (POTS) [542]. As an antioxidant, and probably via other signalling pathways, a particular feature of KYNA is its ability to lower the extent of ischemia-reperfusion injury [517].
Another accompaniment of such diseases is fibrosis and/or amyloid deposition [543,544]. In some cases, fibrin can adopt an amyloid form, e.g., [540,545,546,547,548], though only rarely is fluorescence staining for amyloid performed [549]. Similarly, the ability of KYNA to lower fibrosis [308,550,551,552] may be relevant in this context.
Table 6. Some peripheral disorders in which KYNA has been reported to be protective in mammalian systems.
Table 6. Some peripheral disorders in which KYNA has been reported to be protective in mammalian systems.
Organ/Tissue/DiseaseCommentsSelected References
Alimentary canalProtects vs. stress ulcers in rats[553,554]
Cardiovascular disordersReview[555]
Diabetes, type 2Protective of glomerular filtration rate and against end-stage kidney disease in type 2 diabetes[556]
FibrosisProtective vs. fibrotic injury after surgery[550,552]
HeartProtection against ischemia-reperfusion injury[557,558]
KidneyProtective of glomerular filtration rate and against end-stage kidney disease in type 2 diabetes[556]
Improved kidney function in spontaneously hypertensive and normotensive rats[559]
LiverLevels raised in and protective against hexafluoropropylene oxide dimer acid (HFPO-DA) challenge in mice[560]
Protection vs. nonalcoholic fatty liver disease at very high concentrations[561]
LungProtective in an acute lung injury model [562]
Multi-organProtection against heatstroke by multiple mechanisms, including an anti-apoptotic effect[563]
Pancreatitis (acute)Rat study. Significantly protective at 300 mg/kg[564]
Retinal gangliaProtective against ischemia-reperfusion injury in mice[565]
SepsisProtection vs. neutrophil activation and mitochondrial dysfunction in rats[566]
Active at high doses against LPS-induced inflammation/death in mice[567]
StrokeAssociated with a lower level of risk (but probably also confounded with kynurenine); also protective[568,569,570]
Vascular inflammationProtective[571]
Wound healing and scarring Protective, by largely unknown mechanisms.[550,552,572,573,574]

14.3. Reported Receptors

A number of receptors for KYNA have been detected via ligand binding, albeit often using concentrations far in excess of those measured in vivo. These have recently been reviewed by Turska and colleagues [308], on which Table 7 is partly based. While the data are clear that KYNA can be active at the N-methyl D-aspartate (NMDA) receptor (antagonist), the aryl hydrocarbon receptor (AhR) (agonist), and the GPR35 receptor (agonist), the biological relevance of this awaits an improved understanding of local concentrations of KYNA and other ligands. Meanwhile, the importance of Table 7 for present purposes is more in showing the broad absence of untoward effects at these receptors even when applied concentrations are high.

15. Role of KYNA in Protecting against Ischemia-Reperfusion Injury

Quite a number of the papers mentioned in the above tables highlight a protective role for KYNA in ischemia-reperfusion (I-R) injury, possibly during the hypoxia phase [604]. This I-R injury is a well-known accompaniment [605] in acute circumstances such as stroke [606,607], myocardial infarctions [608,609,610], and organ transplantation [611,612,613,614], as well as in experimental models (e.g., [517,615]). It has recently been recognized as occurring in more chronic circumstances [413] such as long COVID, and as such it is worth highlighting. It occurs when, during a period of hypoxia, commonly caused by ischemia, mitochondria become over-reduced, such that when O2 is readmitted (‘perfusion’), it is reduced not with four electrons as normal (to water) by cytochrome oxidase but to peroxide and superoxide by complexes III and I, respectively (Figure 8). These ‘reactive oxygen species’ can react catalytically with unliganded iron molecules (in the Fenton and Haber–Weiss reactions) to produce the especially damaging hydroxyl radical OH [414], leading to cell death, and the further release of unliganded iron [324], accounting for a number of the symptoms that accompany chronic, inflammatory diseases [410,413]. We consider that analysis of the effects of KYNA on ROS levels and their dynamics is an important and understudied area.

16. Other Factors Known to Affect KYNA Levels

As well as adding KYNA directly, other treatments have been found to increase its level. Many of these are seen as beneficial to the host, though the extent to which KYNA contributes is essentially unknown. We exclude those in which known KYNA precursors are simply added explicitly. Table 8 indicates some.

17. Other Effects of KYNA

Conversely, a number of papers have studied the effects of KYNA addition on different biochemical pathways; although not exhaustive, the point is to show that they are manifold, and they are summarized in Table 9.

18. Safety

Turska and colleagues provide an excellent review [308] of the possibility of enriching foodstuffs with KYNA. By and large, however, human safety studies involving dosing with substantial amounts of pure exogenous KYNA per se have largely not been performed [307], though KYNA (as chestnut honey) was given to human volunteers with no ill effects [306] while a 6 g tryptophan challenge increased KYNA levels more than 130-fold, again without seeming ill-effects, albeit some effects on cerebral blood flow in healthy controls but not in those with schizophrenia-related disorders [664] (rather implying the irrelevance of KYNA here).
Some studies in rodents have added KYNA at massive doses (well over 100 mg/kg, getting into the millimolar range in serum/plasma), seemingly without ill effects (indeed sometimes with protective effects). Table 10 lists some.
We note that Hiratsuka and colleagues [672] gave young female participants up to 5000 mg tryptophan per day (~100 mg/kg based on the stated BMI levels) with no adverse effects; the amount of KYNA that was formed is unknown, though approximately 100 μmol/d (~19 mg) could be found in the urine [672], the value plateauing after ~7 d [673]. Similarly, Al-Karagholi and colleagues [674] dosed volunteers with 5 mg L-kynurenine/kg (say 350 mg total) without apparent ill-effect, while Jauch et al. [675] administered L-kynurenine to rhesus monkeys at doses up to 200 mg/kg, with serum KYNA levels reaching ~25 μM within 10 min, declining to 2.8 μM after 4 h, again without apparent ill effects. Rentschler et al. dosed rats with L-kynurenine at 100 mg/kg, as well as a kynurenine aminotransferase inhibitor at 30 mg/kg; given that on average every marketed drug interacts with at least six known targets [676] (see later section on docking), it is hard to interpret mechanisms during such experiments in which pharmaceutical drugs are added at these kinds of concentration. Note specifically that a hit rate of 1% or more is common in small molecule screens using drug concentrations of just 1–10 μm in individual phenotypic assays (e.g., [258]). L-4-chlorokynurenine is under study as a transportable substrate [677] that can be converted into 7-kynurenic acid, an NMDA receptor antagonist; as part of such studies, doses of L-4-chlorokynurenine of over 1 g per day were well tolerated [678,679]. The main point here is that while any ‘targets’ may remain unknown, there do not seem to be safety issues with these quite substantial doses.

19. Possible Risks

Thus, while the safety profile of KYNA does not yet seem to have been looked at in real detail, we recognize that everyone is different [680], and that there is a tendency for promising studies to become less so over time [681,682,683], not least since possibly ‘occasional’ adverse events are more likely to occur as the populations assessed become larger. We do also note some studies in which KYNA induced possibly undesirable effects. While far less common than those in which it has been shown to be a cytoprotectant, it is appropriate to list some of them (Table 11, see also Table 4). The reasons for such effects are also not really well understood, i.e., whether these are causative or they are essentially downstream responses. As mentioned, the biggest problem with most such studies is that they add tryptophan or L-kynurenine, which can themselves (or their other metabolites) lead to many other undocumented and important changes in host biochemistry. Schizophrenia seems the most common, and as noted, the evidence is at best equivocal as to whether changes in the level of KYNA are a cause or an effect or simply an accompaniment.

20. Regulations for Food Supplements and Nutraceuticals

The use of nutraceuticals varies largely from country to country and, like their use, the legislation around such substances vary widely, lack harmonization and are continuously evolving. Comprehensively, some countries as Canada follow stringent regulations, whereas, in some others, as the United States, well-structured and adequate regulations for nutraceuticals are largely absent [685,686]. Regulations for nutraceuticals in different countries of the world are reviewed in [687,688,689].
In the Unites States, nutraceuticals are included in a category called “dietary supplements”, which is regulated under the Dietary Supplement Health and Education Act of 1994 (DSHEA) through the Food and Drug Administration (FDA). The DSHEA states that a dietary supplement must be intended for ingestion and cannot be designated for other use. However, unlike the stricter FDA’s regulations for drugs, the regulations for dietary supplements are more flexible. The manufacturers and distributors are responsible for nutraceutical safety evaluations and correct labelling, while the FDA limits its role to take action against a product that does not follow the requirements of the DSHEA purely after it reaches the market.
All the dietary supplements that were not marketed before the DSHEA are considered and termed “new dietary ingredient”, and while there is no official list of those that were marketed before the Act, the manufacturer is responsible for determining if a substance is a “new dietary ingredient”. In the case of a “new dietary ingredient”, the manufacturer has to notify the FDA about the ingredient at least 75 days before the product goes into the market; providing information (i.e., any citation to published articles) regarding the expected safety of the new ingredient.
On the other hand, the European Union (EU) has more strict regulations for nutraceuticals completed by the European Food Safety Authority (EFSA), which made a comprehensive assessment of substances that are allowed as a source of such molecules in the EU, including safety of the nutrient source, intake levels and bioavailability. According to the EU General Food Law Regulation (EC) No 178/2002, nutraceuticals are considered foodstuffs, therefore regulated as such. The Directive 2002/46/EC established consensus lists of the vitamins and minerals that can be used for manufacturing food supplements and the labelling requirements for these products, while the use of substances other than vitamins or minerals is subject to the laws prevailing in the different Member States.
In case of foodstuffs not consumed in the EU before 1997, the EFSA requires an application for authorization of a “novel food”, including detailed information about composition, intended use, and safety data before the ingredient is released on the market and it can only be commercialized once authorized. Additionally, according to the Regulation (EC) No 2015/2283, EFSA should provide a scientific opinion on a substance’s safety when it undergoes an application for “novel foods”.

21. Analytics

Since the coining of the term in 1998 [690], metabolomics studies are well into their third decade [691], where the analysis of small molecules such as KYNA is now dominated by methods combining gas or liquid chromatography with mass spectrometry (e.g., [692,693]). As an aromatic amino acid that ionizes reasonably well, a variety of such analytical methods have indeed been developed (e.g., [307,339,380,381,694,695,696,697,698,699]). In addition, as a redox-active fluorophore (whose fluorescence is enhanced by zinc ions [40,700]), KYNA may also be detected by electrochemistry [35,356] and optically [352,700,701,702,703,704,705,706,707,708,709,710], or (as any organic molecule) by vibrational spectroscopy [711].

22. KYNA as a Therapeutic for Chronic Inflammatory Diseases

Many of the diseases mentioned above share similar properties, including in particular that they are chronic and accompanied by inflammation (an interesting recent suggestion around the latter based on mitochondrial antipathy to their cellular host is worth flagging [712]). Thus, any nutraceuticals that might be able to tackle inflammation would be of value, and we have set out here the evidence that leads us to suppose that KYNA might be one. This said, there are seemingly some chronic inflammatory diseases, such as rheumatoid disease [713], that seem not to have major changes in KYNA.

23. Use of KYNA as an Antioxidant in Processed Foodstuffs

Although ergothioneine has been used successfully in this way as an antioxidant, e.g., in seafoods [714,715,716], we are not aware of any attempts to use KYNA in this way. We note, however, the important work on the suggestions of fortifying artificial baby milk with it [359] on the grounds that its levels are significantly lower than those of natural human milk.

24. KYNA in the Feed of Racing Animals

It is implicit that if KYNA is a nutraceutical, it may have value for elite athletes as well as for the ‘normal’ population. Purified KYNA is unlikely to be economically competitive as an additive in the feed of food animals, but its addition may well be worthwhile for those involved in horse or camel racing. However, the analysis of tryptophan metabolism in such animals is in its infancy [717].

25. Use of KYNA in Cosmetics

‘Cosmeceuticals’ are nutraceuticals that are marketed for their cosmetic benefits (e.g., [718,719,720,721,722,723,724]). Because significant skin damage is caused by UV-mediated ROS production [725,726,727,728], it is reasonable—much as with ergothioneine [18]—that KYNA might be useful as a cosmeceutical, and it can both be formulated in creams and taken up into the body [551]. As with ergothioneine, it is possible that its relative unavailability is holding back such uses here, though we note that it is also a photosensitizer. Its value as a topical treatment in inhibiting scarring has, however, been demonstrated [550,573], and Aryl hydrocarbon receptor agonists (of which KYNA is one, and including the FDA-approved tapinarof [729,730]) have been shown to have benefits in both psoriasis and atopic dermatitis [731,732,733,734,735]. Thus, KYNA would seem well worth exploring as a cosmeceutical ingredient.

26. Role of KYNA as a Cofactor

While KYNA is clearly capable of acting as an antioxidant directly (as can ergothioneine), most small molecules (including vitamins) interact with proteins of various kinds, and many more than we usually credit (e.g., [736,737,738,739,740]. In addition, the relatively low concentrations of KYNA in humans also imply a more regulatory role that can lead to genetic induction or repression and thus amplification of their signal. In this vein, as our ‘index’ antioxidant nutraceutical, ergothioneine acts in part via the redox-active transcription factor Nrf2 (e.g., [741,742,743,744,745,746]). Studies in KYNA are far behind, but a tantalizing report [747] shows that chestnut honey—the foodstuff containing by far the largest amount of KYNA (Table 3)—can exert protective effects via Nrf2 on LPS-treated macrophages and indomethacin-treated gastric mucosa. Indeed, high levels of KYNA can induce Nrf2 synthesis [94] and prevent the induction of reactive oxygen species [95] and other changes [748] caused by quinolinic acid. While the aryl hydrocarbon receptor AhR is also a transcription factor, its activation can itself stimulate the activation of Nrf2 [732,733,749], and a variety of known agonists target both ArH and Nrf2 [731,750] and assist with atopic dermatitis and psoriasis (see above), so this is reasonable. On the other hand it is activated by dioxins, with somewhat negative effects [749,751], and has a complex expression profile in certain tumors [752]. Its expression is also affected by NF-κB [752,753], a transcription factor whose activity depends on frequency rather than amplitude [754,755,756], and this is still rarely recorded. Deconvolving the detailed interactions between KYNA and AhR is thus a highly non-trivial process.

27. Cheminformatics of KYNA

One strategy for understanding the biology of a small molecule is to assess how close it is to other endogenous metabolites and, in particular, to marketed drugs, as knowledge of their binding partners or mode of action might give clues to the binding partners of KYNA. One paper shows such an analysis using the RDKit [757]. Pattern encoding and the otherwise precise methods are described in detail elsewhere [310,758,759,760,761]. Only four marketed drugs have a Tanimoto similarity exceeding 0.7, and these are displayed in Figure 9. Interestingly, nalidixic acid is transported in E. coli via the fadL fatty acid transporter [762] (not studied in [763]), so this kind of observation may provide clues. However, since this is not our present focus, we simply set out these data and thoughts to guide future studies.

28. Predicting the KYNA Interactome

Multiple cheminformatic tools now exist for compound–target interaction prediction in silico (e.g., [764,765]) (we ignore those based on generative AI, as they are still in their infancy [766], though this is changing rapidly, and most seek molecules that bind to specified targets, not the other way round as we are interested in here). These tools employ a variety of network-based approaches, machine-learning models, and molecular-docking algorithms to predict the binding of small molecules to target proteins or receptors. As prediction results depend on both the underlying Knowledgebases and the computational approaches applied, it is prudent to examine both the intersection and the compiled results, including with pathway topology and other functional annotation approaches from multiple prediction tools. We have investigated potential KYNA binding targets using three such OpenSource tools with recently updated databases and differing prediction approaches. Specifically, we used PharmMapper [767] that uses a reverse pharmacophore mapping approach and requires a 3D structure (mol2 or sdf format) for ligand input, SwissTargetPrediction [768] that examines 2D and 3D molecular similarity, and SuperPred 3.0 [769] that uses machine learning models for prediction. SwissTargetPrediction and SuperPred take ligand input in a simplified molecular-input line-entry system (SMILES) [770] format. The full dataset is given in Supplementary Spreadsheet S1. Remarkably, only one protein was predicted by all three tools, the thyroid hormone receptor alpha (THA, Figure 10A). This is especially interesting, as previously impaired removal of KYNA from the brain during has been observed in experimental hypothyroidism [625] (Table 8). Twenty-seven predicted targets for KYNA were identified by at least two of these computational compound-target prediction tools, including the aforementioned AhR (Table 12). Other nuclear receptor family members predicted in addition to THA and AhR, included the peroxisome proliferator-activated receptor alpha (PPAR-alpha) and estrogen receptor beta (ER-beta). The latter is notable in the light of a recent report of lower plasma KYNA levels in users of estrogen contraceptives [632] (Table 8). Notably, however, several of those in Table 7 were not picked up using this approach. Also of interest is that KYNA was predicted to bind multiple isomers of the zinc-containing enzyme carbonic anhydrase, which seems at least plausible as tryptophan has been shown by crystallography to bind and activate carbonic anhydrase 2 [771]. Not least of note is the prediction that KYNA was predicted to bind the lymphocyte specific tyrosine kinase (LCK). Critical to T cell signalling, LCK function has been shown by us to be exquisitely sensitive to dietary zinc supply [772], and this too may contribute to the immunological effects caused by KYNA.
As experimentally validated and/or plausible targets were among the proteins predicted by only one tool (e.g., acetylcholine receptors from SuperPred, SLC16A1 from SwissTargetPrediction, albumin from PharmMapper) we performed functional enrichment analyses in the DAVID Knowledgebase [773] on all (n = 455) predicted interactors (Figure 10B). The top functional clusters included carbonic anhydrase activity and multiple neuromodulatory receptor and ion channel activities. Notably, neurotransmitter function and neurodegenerative disease were among the top biological functional clusters associated with the predicted KYNA interactome (Figure 10B).

29. Biotechnological Production

Current laboratory [774,775,776] and commercial production of KYNA is via chemical synthesis, which presently uses some environmentally unpleasant chemicals and has modest yields. However, as with ergothioneine (e.g., [18,777,778,779]), it is possible to product KYNA by fermentation, Studies of the fermentative production of KYNA are relatively limited, however, and are summarized in Table 13.

30. Conclusions and Forward Look

Thanks to advances in scientific knowledge and in public health, human lifespan has been increasing in developed countries since the middle of the 19th century at something like 6 y per 25 y (ungendered and aggregated data for the UK at https://www.statista.com/statistics/1040159/life-expectancy-united-kingdom-all-time/, accessed on 19 August 2024), and its variation between individuals and countries has also decreased [786]. However, the healthspan, the period in which one is free of significant ill-health, has not matched it [787]. It is widely recognized (e.g., [788,789,790]) that diet has a significant role in bringing the healthspan closer to the lifespan. In particular, nutraceuticals, molecules that influence health positively and that might be part of or added to a ‘healthy diet’, are seen as a contributor. Ergothioneine (ERG), a potent and effective antioxidant, seems to be one [1,18,29,508,791,792], with experimental evidence and ongoing studies for this continuing to emerge (e.g., [320,793]), and we here make the case that KYNA is another. It is instructive to provide a comparison of the two molecules, and Table 14 does so.
Overall, the evidence that KYNA may indeed be a worthwhile nutraceutical that might be added exogenously involves at least the following:
  • Many examples in which its exogenous addition seems to offer benefits of health or of protection against disease
  • Evidence that its concentration is relatively low in normal populations
  • Safety evidence to the effect that there do not seem to be examples in which hyperactive alleles of KAT enzymes lead to overt disease, and that exogenous KYNA cannot realistically ‘go back’ to L-kynurenine
  • Evidence that it is more or less readily bioavailable for entering plasma from the diet rather than simply being produced by compounds such as tryptophan and L-kynurenine that are more easily transported but that can lead to other, potentially toxic molecules.
This said, not least by comparison with ergothioneine, there are considerable gaps in our knowledge of its biology. In the case of ergothioneine, there has been a massive upsurge in interest in the last 20 years, and as pointed out by Halliwell and Cheah [792], “a key factor was the discovery that an organic cation transporter, OCTN1, is responsible for uptake of “ergothioneine” from the gastrointestinal tract and for its distribution to tissues in the bodies of humans and other animals” (see [802,803]). Given the evidence for ‘missing’ transporters that we present above, a similar trajectory seems plausible for KYNA.
From the perspective of nutrition, the number of foodstuffs for which KYNA contents have been measured by multiple laboratories using modern, quantitative methods is rather limited, and such studies demand extension.
Other glaring gaps in our knowledge involve pharmacokinetic studies of KYNA uptake, distribution, metabolism and excretion in both humans and laboratory animals (for ergothioneine, see e.g., [319]), the effects of KYNA supplementation on measures of health such as antioxidant status and indeed longevity, and other studies manipulating KYNA as an independent variable to establish suitable nutraceutical dosing levels. We may also be sure that it interacts with other proteins whose identity has not yet been discovered, not least for some of the other targets that were predicted by in silico docking (Table 12), where modern proteomics approaches to attack such questions are available [740,811,812].
The biggest issue with many studies where, for example, tryptophan or L-kynurenine was added, is that they often infer effects of KYNA that are equally plausibly due to changes in other metabolites of the kynurenine pathway or elsewhere that were not in fact measured. In a sense, this is one of the great strengths of KYNA as a candidate nutraceutical, as it can be added without being expected to affect upstream metabolites significantly, at least directly. This offers particular levels of safety.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25169082/s1.

Author Contributions

Conceptualization, D.B.K.; methodology, all authors; validation, all authors; data curation, all authors.; writing—original draft preparation, D.B.K.; writing—review and editing, all authors; visualization, all authors; supervision, D.B.K. and J.B.M.; project administration, D.B.K.; funding acquisition, D.B.K. All authors have read and agreed to the published version of the manuscript.

Funding

The Balvi Foundation (grant 18) and the Novo Nordisk Foundation (grant NNF20CC0035580) for financial support.

Acknowledgments

We thank Pedro Mendes (UConn Health) for running the SBML-based ODE model referred to in the text. The APC was funded by the journal.

Conflicts of Interest

The authors report no conflicts of interest.

References

  1. Ames, B.N. Prolonging healthy aging: Longevity vitamins and proteins. Proc. Natl. Acad. Sci. USA 2018, 115, 10836–10844. [Google Scholar] [CrossRef]
  2. Bahadoran, Z.; Mirmiran, P.; Azizi, F. Dietary polyphenols as potential nutraceuticals in management of diabetes: A review. J. Diabetes Metab. Disord. 2013, 12, 43. [Google Scholar] [CrossRef] [PubMed]
  3. Ogle, W.O.; Speisman, R.B.; Ormerod, B.K. Potential of treating age-related depression and cognitive decline with nutraceutical approaches: A mini-review. Gerontology 2013, 59, 23–31. [Google Scholar] [CrossRef] [PubMed]
  4. Ragle, R.L.; Sawitzke, A.D. Nutraceuticals in the management of osteoarthritis: A critical review. Drugs Aging 2012, 29, 717–731. [Google Scholar] [CrossRef]
  5. Chauhan, B.; Kumar, G.; Kalam, N.; Ansari, S.H. Current concepts and prospects of herbal nutraceutical: A review. J. Adv. Pharm. Technol. Res. 2013, 4, 4–8. [Google Scholar] [CrossRef]
  6. Borghi, C.; Cicero, A.F. Nutraceuticals with a clinically detectable blood pressure-lowering effect: A review of available randomized clinical trials and their meta-analyses. Br. J. Clin. Pharmacol. 2017, 83, 163–171. [Google Scholar] [CrossRef]
  7. Aruoma, O.I.; Coles, L.S.; Landes, B.; Repine, J.E. Functional benefits of ergothioneine and fruit- and vegetable-derived nutraceuticals: Overview of the supplemental issue contents. Prev. Med. 2012, 54, S4–S8. [Google Scholar] [CrossRef]
  8. Rathore, H.; Prasad, S.; Sharma, S. Mushroom nutraceuticals for improved nutrition and better human health: A review. PharmaNutrition 2017, 5, 35–46. [Google Scholar]
  9. Cencic, A.; Chingwaru, W. The role of functional foods, nutraceuticals, and food supplements in intestinal health. Nutrients 2010, 2, 611–625. [Google Scholar] [CrossRef] [PubMed]
  10. Espín, J.C.; García-Conesa, M.T.; Tomás-Barberán, F.A. Nutraceuticals: Facts and fiction. Phytochemistry 2007, 68, 2986–3008. [Google Scholar] [CrossRef]
  11. Sharif, M.K.; Khalid, R. Nutraceuticals: Myths Versus Realities. In Therapeut Foods; Elsevier: Amsterdam, The Netherlands, 2018; pp. 3–21. [Google Scholar]
  12. Singh, S.; Razak, M.A.; Sangam, S.R.; Viswanath, B.; Begum, P.S.; Rajagopal, S. The Impact of Functional Food and Nutraceuticals in Health. In Therapeut Foods; Academic Press: Cambridge, MA, USA, 2018; pp. 23–47. [Google Scholar]
  13. Spindler, S.R.; Mote, P.L.; Flegal, J.M. Lifespan effects of simple and complex nutraceutical combinations fed isocalorically to mice. Age 2014, 36, 705–718. [Google Scholar] [CrossRef]
  14. D’Cunha, N.M.; Georgousopoulou, E.N.; Dadigamuwage, L.; Kellett, J.; Panagiotakos, D.B.; Thomas, J.; McKune, A.J.; Mellor, D.D.; Naumovski, N. Effect of long-term nutraceutical and dietary supplement use on cognition in the elderly: A 10-year systematic review of randomised controlled trials. Br. J. Nutr. 2018, 119, 280–298. [Google Scholar] [CrossRef]
  15. Hopper, I.; Connell, C.; Briffa, T.; De Pasquale, C.G.; Driscoll, A.; Kistler, P.M.; Macdonald, P.S.; Sindone, A.; Thomas, L.; Atherton, J.J. Nutraceuticals in Patients With Heart Failure: A Systematic Review. J. Card. Fail. 2020, 26, 166–179. [Google Scholar] [CrossRef]
  16. Ilari, S.; Proietti, S.; Russo, P.; Malafoglia, V.; Gliozzi, M.; Maiuolo, J.; Oppedisano, F.; Palma, E.; Tomino, C.; Fini, M.; et al. A Systematic Review and Meta-Analysis on the Role of Nutraceuticals in the Management of Neuropathic Pain in In Vivo Studies. Antioxidants 2022, 11, 2361. [Google Scholar] [CrossRef]
  17. Bartel, I.; Mandryk, I.; Horbańczuk, J.O.; Wierzbicka, A.; Koszarska, M. Nutraceutical Properties of Syringic Acid in Civilization Diseases-Review. Nutrients 2023, 16, 10. [Google Scholar] [CrossRef]
  18. Borodina, I.; Kenny, L.C.; McCarthy, C.M.; Paramasivan, K.; Pretorius, R.; Roberts, T.J.; van der Hoek, S.A.; Kell, D.B. The biology of ergothioneine, an antioxidant nutraceutical. Nutr. Res. Rev. 2020, 33, 190–217. [Google Scholar] [CrossRef]
  19. Matera, R.; Lucchi, E.; Valgimigli, L. Plant Essential Oils as Healthy Functional Ingredients of Nutraceuticals and Diet Supplements: A Review. Molecules 2023, 28, 901. [Google Scholar] [CrossRef] [PubMed]
  20. Lippi, L.; Uberti, F.; Folli, A.; Turco, A.; Curci, C.; d’Abrosca, F.; de Sire, A.; Invernizzi, M. Impact of nutraceuticals and dietary supplements on mitochondria modifications in healthy aging: A systematic review of randomized controlled trials. Aging Clin. Exp. Res. 2022, 34, 2659–2674. [Google Scholar] [CrossRef] [PubMed]
  21. Allaqaband, S.; Dar, A.H.; Patel, U.; Kumar, N.; Nayik, G.A.; Khan, S.A.; Ansari, M.J.; Alabdallah, N.M.; Kumar, P.; Pandey, V.K.; et al. Utilization of Fruit Seed-Based Bioactive Compounds for Formulating the Nutraceuticals and Functional Food: A Review. Front. Nutr. 2022, 9, 902554. [Google Scholar] [CrossRef]
  22. Hosseini, S.F.; Rezaei, M.; McClements, D.J. Bioactive functional ingredients from aquatic origin: A review of recent progress in marine-derived nutraceuticals. Crit. Rev. Food Sci. Nutr. 2022, 62, 1242–1269. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, Y.; Xu, S.; Tang, L.; Gong, J.; Su, D.; Yang, H. Piperine as a Potential Nutraceutical Agent for Managing Diabetes and Its Complications: A Literature Review. J. Med. Food 2023, 26, 693–704. [Google Scholar] [CrossRef]
  24. Yılmaz, C.; Gökmen, V. Perspective on the Formation, Analysis, and Health Effects of Neuroactive Compounds in Foods. J. Agric. Food Chem. 2021, 69, 13364–13372. [Google Scholar] [CrossRef] [PubMed]
  25. Puri, V.; Nagpal, M.; Singh, I.; Singh, M.; Dhingra, G.A.; Huanbutta, K.; Dheer, D.; Sharma, A.; Sangnim, T. A Comprehensive Review on Nutraceuticals: Therapy Support and Formulation Challenges. Nutrients 2022, 14, 4637. [Google Scholar] [CrossRef] [PubMed]
  26. Temple, N.J. A rational definition for functional foods: A perspective. Front. Nutr. 2022, 9, 957516. [Google Scholar] [CrossRef]
  27. Ruscica, M.; Penson, P.E.; Ferri, N.; Sirtori, C.R.; Pirro, M.; Mancini, G.B.J.; Sattar, N.; Toth, P.P.; Sahebkar, A.; Lavie, C.J.; et al. Impact of nutraceuticals on markers of systemic inflammation: Potential relevance to cardiovascular diseases—A position paper from the International Lipid Expert Panel (ILEP). Prog. Cardiovasc. Dis. 2021, 67, 40–52. [Google Scholar] [CrossRef] [PubMed]
  28. Rizzo, M.; Colletti, A.; Penson, P.E.; Katsiki, N.; Mikhailidis, D.P.; Toth, P.P.; Gouni-Berthold, I.; Mancini, J.; Marais, D.; Moriarty, P.; et al. Nutraceutical approaches to non-alcoholic fatty liver disease (NAFLD): A position paper from the International Lipid Expert Panel (ILEP). Pharmacol. Res. 2023, 189, 106679. [Google Scholar] [CrossRef] [PubMed]
  29. Tian, X.; Thorne, J.L.; Moore, J.B. Ergothioneine: An underrecognised dietary micronutrient required for healthy ageing? Br. J. Nutr. 2023, 129, 104–114. [Google Scholar] [CrossRef] [PubMed]
  30. Liebig, J. Über Kynurensäure. Ann. Chem. 1853, 86, 125–126. [Google Scholar]
  31. Pileni, M.P.; Giraud, M.; Santus, R. Kynurenic acid: I. Spectroscopic properties. Photochem. Photobiol. 1979, 30, 251–256. [Google Scholar]
  32. Zelentsova, E.A.; Sherin, P.S.; Snytnikova, O.A.; Kaptein, R.; Vauthey, E.; Tsentalovich, Y.P. Photochemistry of aqueous solutions of kynurenic acid and kynurenine yellow. Photochem. Photobiol. Sci. 2013, 12, 546–558. [Google Scholar] [CrossRef] [PubMed]
  33. Bellmaine, S.; Schnellbaecher, A.; Zimmer, A. Reactivity and degradation products of tryptophan in solution and proteins. Free Radic. Biol. Med. 2020, 160, 696–718. [Google Scholar] [CrossRef] [PubMed]
  34. Pogoda, D.; Janczak, J.; Pawlak, S.; Zaworotko, M.; Videnova-Adrabinska, V. Tautomeric polymorphism of the neuroactive inhibitor kynurenic acid. Acta Crystallogr. C Struct. Chem. 2019, 75, 793–805. [Google Scholar] [CrossRef] [PubMed]
  35. Sadok, I.; Staniszewska, M. Electrochemical Determination of Kynurenine Pathway Metabolites-Challenges and Perspectives. Sensors 2021, 21, 7152. [Google Scholar] [CrossRef] [PubMed]
  36. Turski, M.P.; Turska, M.; Zgrajka, W.; Kuc, D.; Turski, W.A. Presence of kynurenic acid in food and honeybee products. Amino Acids 2009, 36, 75–80. [Google Scholar] [CrossRef] [PubMed]
  37. Kita, A.; Kołodziejczyk, M.; Michalska-Ciechanowska, A.; Brzezowska, J.; Wicha-Komsta, K.; Turski, W. The Effect of Thermal Treatment on Selected Properties and Content of Biologically Active Compounds in Potato Crisps. Appl. Sci. 2022, 12, 555. [Google Scholar] [CrossRef]
  38. Pileni, M.P.; Giraud, M.; Santus, R. Kynurenic acid: II. Photosensitizing properties. Photochem. Photobiol. 1979, 30, 257–261. [Google Scholar]
  39. Samanta, A.; Guchhait, N.; Bhattacharya, S.C. Photophysical aspects of biological photosensitizer Kynurenic acid from the perspective of experimental and quantum chemical study. Spectrochim. Acta 2014, 129, 457–465. [Google Scholar] [CrossRef]
  40. Diana, R.; Panunzi, B. The Role of Zinc(II) Ion in Fluorescence Tuning of Tridentate Pincers: A Review. Molecules 2020, 25, 4984. [Google Scholar] [CrossRef]
  41. Kubicova, L.; Hadacek, F.; Bachmann, G.; Weckwerth, W.; Chobot, V. Coordination Complex Formation and Redox Properties of Kynurenic and Xanthurenic Acid Can Affect Brain Tissue Homeodynamics. Antioxidants 2019, 8, 476. [Google Scholar] [CrossRef] [PubMed]
  42. Fukushima, T.; Mitsuhashi, S.; Tomiya, M.; Kawai, J.; Hashimoto, K.; Toyo’oka, T. Determination of rat brain kynurenic acid by column-switching HPLC with fluorescence detection. Biomed. Chromatogr. 2007, 21, 514–519. [Google Scholar] [CrossRef]
  43. Qi, X.; Fu, K.; Yue, M.; Shou, N.; Yuan, X.; Chen, X.; He, C.; Yang, Y.; Shi, Z. Kynurenic acid mediates bacteria-algae consortium in resisting environmental cadmium toxicity. J. Hazard. Mater. 2023, 444, 130397. [Google Scholar] [CrossRef] [PubMed]
  44. Richard, D.M.; Dawes, M.A.; Mathias, C.W.; Acheson, A.; Hill-Kapturczak, N.; Dougherty, D.M. L-Tryptophan: Basic Metabolic Functions, Behavioral Research and Therapeutic Indications. Int. J. Tryptophan Res. 2009, 2, 45–60. [Google Scholar] [CrossRef] [PubMed]
  45. Muhamadali, H.; Winder, C.L.; Dunn, W.B.; Goodacre, R. Unlocking the secrets of the microbiome: Exploring the dynamic microbial interplay with humans through metabolomics and their manipulation for synthetic biology applications. Biochem. J. 2023, 480, 891–908. [Google Scholar] [CrossRef]
  46. McCann, J.R.; Rawls, J.F. Essential Amino Acid Metabolites as Chemical Mediators of Host-Microbe Interaction in the Gut. Annu. Rev. Microbiol. 2023, 77, 479–497. [Google Scholar] [CrossRef]
  47. Miyamoto, K.; Sujino, T.; Kanai, T. The tryptophan metabolic pathway of the microbiome and host cells in health and disease. Int. Immunol. 2024. [Google Scholar] [CrossRef]
  48. Schwarcz, R.; Foo, A.; Sathyasaikumar, K.V.; Notarangelo, F.M. The Probiotic Lactobacillus reuteri Preferentially Synthesizes Kynurenic Acid from Kynurenine. Int. J. Mol. Sci. 2024, 25, 3679. [Google Scholar] [CrossRef]
  49. Zhu, F.; Guo, R.; Wang, W.; Ju, Y.; Wang, Q.; Ma, Q.; Sun, Q.; Fan, Y.; Xie, Y.; Yang, Z.; et al. Transplantation of microbiota from drug-free patients with schizophrenia causes schizophrenia-like abnormal behaviors and dysregulated kynurenine metabolism in mice. Mol. Psychiatry 2020, 25, 2905–2918. [Google Scholar] [CrossRef]
  50. Kennedy, P.J.; Cryan, J.F.; Dinan, T.G.; Clarke, G. Kynurenine pathway metabolism and the microbiota-gut-brain axis. Neuropharmacology 2017, 112, 399–412. [Google Scholar] [CrossRef] [PubMed]
  51. Dehhaghi, M.; Kazemi Shariat Panahi, H.; Guillemin, G.J. Microorganisms, Tryptophan Metabolism, and Kynurenine Pathway: A Complex Interconnected Loop Influencing Human Health Status. Int. J. Tryptophan Res. 2019, 12, 1178646919852996. [Google Scholar] [CrossRef]
  52. Gao, K.; Mu, C.L.; Farzi, A.; Zhu, W.Y. Tryptophan Metabolism: A Link Between the Gut Microbiota and Brain. Adv. Nutr. 2020, 11, 709–723. [Google Scholar] [CrossRef]
  53. Chen, L.M.; Bao, C.H.; Wu, Y.; Liang, S.H.; Wang, D.; Wu, L.Y.; Huang, Y.; Liu, H.R.; Wu, H.G. Tryptophan-kynurenine metabolism: A link between the gut and brain for depression in inflammatory bowel disease. J. Neuroinflammation 2021, 18, 135. [Google Scholar] [CrossRef]
  54. Nagy-Grócz, G.; Spekker, E.; Vécsei, L. Kynurenines, Neuronal Excitotoxicity, and Mitochondrial Oxidative Stress: Role of the Intestinal Flora. Int. J. Mol. Sci. 2024, 25, 1698. [Google Scholar] [CrossRef] [PubMed]
  55. Yılmaz, C.; Gökmen, V. Neuroactive compounds in foods: Occurrence, mechanism and potential health effects. Food Res. Int. 2020, 128, 108744. [Google Scholar] [CrossRef]
  56. Kaur, H.; Bose, C.; Mande, S.S. Tryptophan Metabolism by Gut Microbiome and Gut-Brain-Axis: An in silico Analysis. Front. Neurosci. 2019, 13, 1365. [Google Scholar] [CrossRef]
  57. Fila, M.; Chojnacki, J.; Pawlowska, E.; Szczepanska, J.; Chojnacki, C.; Blasiak, J. Kynurenine Pathway of Tryptophan Metabolism in Migraine and Functional Gastrointestinal Disorders. Int. J. Mol. Sci. 2021, 22, 10134. [Google Scholar] [CrossRef]
  58. Malan-Muller, S.; Valles-Colomer, M.; Raes, J.; Lowry, C.A.; Seedat, S.; Hemmings, S.M.J. The Gut Microbiome and Mental Health: Implications for Anxiety- and Trauma-Related Disorders. OMICS 2018, 22, 90–107. [Google Scholar] [CrossRef]
  59. Rust, C.; Malan-Muller, S.; van den Heuvel, L.L.; Tonge, D.; Seedat, S.; Pretorius, E.; Hemmings, S.M.J. Platelets bridging the gap between gut dysbiosis and neuroinflammation in stress-linked disorders: A narrative review. J. Neuroimmunol. 2023, 382, 578155. [Google Scholar] [CrossRef] [PubMed]
  60. Chen, C.Y.; Wang, Y.F.; Lei, L.; Zhang, Y. Impacts of microbiota and its metabolites through gut-brain axis on pathophysiology of major depressive disorder. Life Sci. 2024, 351, 122815. [Google Scholar] [CrossRef]
  61. Ramos-Chávez, L.A.; Lugo Huitrón, R.; González Esquivel, D.; Pineda, B.; Ríos, C.; Silva-Adaya, D.; Sánchez-Chapul, L.; Roldán-Roldán, G.; Pérez de la Cruz, V. Relevance of Alternative Routes of Kynurenic Acid Production in the Brain. Oxid. Med. Cell Longev. 2018, 2018, 5272741. [Google Scholar] [CrossRef]
  62. Stone, T.W.; Darlington, L.G. Endogenous kynurenines as targets for drug discovery and development. Nat. Rev. Drug Discov. 2002, 1, 609–620. [Google Scholar] [CrossRef]
  63. Badawy, A.A. Hypothesis kynurenic and quinolinic acids: The main players of the kynurenine pathway and opponents in inflammatory disease. Med. Hypotheses 2018, 118, 129–138. [Google Scholar] [CrossRef]
  64. Badawy, A.A. Kynurenine Pathway of Tryptophan Metabolism: Regulatory and Functional Aspects. Int. J. Tryptophan Res. 2017, 10, 1178646917691938. [Google Scholar] [CrossRef] [PubMed]
  65. Stavrum, A.K.; Heiland, I.; Schuster, S.; Puntervoll, P.; Ziegler, M. Model of tryptophan metabolism, readily scalable using tissue-specific gene expression data. J. Biol. Chem. 2013, 288, 34555–34566. [Google Scholar] [CrossRef] [PubMed]
  66. Rossi, F.; Miggiano, R.; Ferraris, D.M.; Rizzi, M. The Synthesis of Kynurenic Acid in Mammals: An Updated Kynurenine Aminotransferase Structural KATalogue. Front. Mol. Biosci. 2019, 6, 7. [Google Scholar] [CrossRef]
  67. Kolodziej, L.R.; Paleolog, E.M.; Williams, R.O. Kynurenine metabolism in health and disease. Amino Acids 2011, 41, 1173–1183. [Google Scholar] [CrossRef]
  68. Proietti, E.; Rossini, S.; Grohmann, U.; Mondanelli, G. Polyamines and Kynurenines at the Intersection of Immune Modulation. Trends Immunol. 2020, 41, 1037–1050. [Google Scholar] [CrossRef] [PubMed]
  69. Biernacki, T.; Sandi, D.; Bencsik, K.; Vécsei, L. Kynurenines in the Pathogenesis of Multiple Sclerosis: Therapeutic Perspectives. Cells 2020, 9, 1564. [Google Scholar] [CrossRef]
  70. Guillemin, G.J.; Smythe, G.; Takikawa, O.; Brew, B.J. Expression of indoleamine 2,3-dioxygenase and production of quinolinic acid by human microglia, astrocytes, and neurons. Glia 2005, 49, 15–23. [Google Scholar] [CrossRef]
  71. Parrott, J.M.; O’Connor, J.C. Kynurenine 3-Monooxygenase: An Influential Mediator of Neuropathology. Front. Psychiatry 2015, 6, 116. [Google Scholar] [CrossRef]
  72. Vécsei, L.; Szalárdy, L.; Fülöp, F.; Toldi, J. Kynurenines in the CNS: Recent advances and new questions. Nat. Rev. Drug Discov. 2013, 12, 64–82. [Google Scholar] [CrossRef]
  73. Davey, H.M.; Kell, D.B. Flow cytometry and cell sorting of heterogeneous microbial populations: The importance of single-cell analysis. Microbiol. Rev. 1996, 60, 641–696. [Google Scholar] [PubMed]
  74. Lundberg, E.; Uhlén, M. Creation of an antibody-based subcellular protein atlas. Proteomics 2010, 10, 3984–3996. [Google Scholar] [PubMed]
  75. Uhlén, M.; Oksvold, P.; Fagerberg, L.; Lundberg, E.; Jonasson, K.; Forsberg, M.; Zwahlen, M.; Kampf, C.; Wester, K.; Hober, S.; et al. Towards a knowledge-based Human Protein Atlas. Nat. Biotechnol. 2010, 28, 1248–1250. [Google Scholar] [PubMed]
  76. Regev, A.; Teichmann, S.A.; Lander, E.S.; Amit, I.; Benoist, C.; Birney, E.; Bodenmiller, B.; Campbell, P.; Carninci, P.; Clatworthy, M.; et al. The Human Cell Atlas. eLife 2017, 6, e27041. [Google Scholar] [CrossRef]
  77. George, N.; Fexova, S.; Fuentes, A.M.; Madrigal, P.; Bi, Y.; Iqbal, H.; Kumbham, U.; Nolte, N.F.; Zhao, L.; Thanki, A.S.; et al. Expression Atlas update: Insights from sequencing data at both bulk and single cell level. Nucleic Acids Res. 2024, 52, D107–D114. [Google Scholar] [CrossRef] [PubMed]
  78. Lindeboom, R.G.H.; Regev, A.; Teichmann, S.A. Towards a Human Cell Atlas: Taking Notes from the Past. Trends Genet. 2021, 37, 625–630. [Google Scholar] [CrossRef] [PubMed]
  79. Rood, J.E.; Maartens, A.; Hupalowska, A.; Teichmann, S.A.; Regev, A. Impact of the Human Cell Atlas on medicine. Nat. Med. 2022, 28, 2486–2496. [Google Scholar] [CrossRef] [PubMed]
  80. Savitz, J. The kynurenine pathway: A finger in every pie. Mol. Psychiatry 2020, 25, 131–147. [Google Scholar] [CrossRef] [PubMed]
  81. Joisten, N.; Ruas, J.L.; Braidy, N.; Guillemin, G.J.; Zimmer, P. The kynurenine pathway in chronic diseases: A compensatory mechanism or a driving force? Trends Mol. Med. 2021, 27, 946–954. [Google Scholar] [CrossRef] [PubMed]
  82. Zsizsik, B.K.; Hardeland, R. Formation of kynurenic and xanthurenic acids from kynurenine and 3-hydroxykynurenine in the dinoflagellate Lingulodinium polyedrum: Role of a novel, oxidative pathway. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2002, 133, 383–392. [Google Scholar] [CrossRef] [PubMed]
  83. Pabarcus, M.K.; Casida, J.E. Kynurenine formamidase: Determination of primary structure and modeling-based prediction of tertiary structure and catalytic triad. Biochim. Biophys. Acta 2002, 1596, 201–211. [Google Scholar] [CrossRef] [PubMed]
  84. Lee, J.M.; Tan, V.; Lovejoy, D.; Braidy, N.; Rowe, D.B.; Brew, B.J.; Guillemin, G.J. Involvement of quinolinic acid in the neuropathogenesis of amyotrophic lateral sclerosis. Neuropharmacology 2017, 112, 346–364. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, S.; Sakuma, M.; Deora, G.S.; Levy, C.W.; Klausing, A.; Breda, C.; Read, K.D.; Edlin, C.D.; Ross, B.P.; Wright Muelas, M.; et al. A brain-permeable inhibitor of the neurodegenerative disease target kynurenine 3-monooxygenase prevents accumulation of neurotoxic metabolites. Commun. Biol. 2019, 2, 271. [Google Scholar] [CrossRef]
  86. Tóth, F.; Cseh, E.K.; Vécsei, L. Natural Molecules and Neuroprotection: Kynurenic Acid, Pantethine and alpha-Lipoic Acid. Int. J. Mol. Sci. 2021, 22, 403. [Google Scholar] [CrossRef]
  87. Perkins, M.N.; Stone, T.W. An iontophoretic investigation of the actions of convulsant kynurenines and their interaction with the endogenous excitant quinolinic acid. Brain Res. 1982, 247, 184–187. [Google Scholar] [CrossRef] [PubMed]
  88. Foster, A.C.; Vezzani, A.; French, E.D.; Schwarcz, R. Kynurenic acid blocks neurotoxicity and seizures induced in rats by the related brain metabolite quinolinic acid. Neurosci. Lett. 1984, 48, 273–278. [Google Scholar] [CrossRef] [PubMed]
  89. Schwarcz, R.; Bruno, J.P.; Muchowski, P.J.; Wu, H.Q. Kynurenines in the mammalian brain: When physiology meets pathology. Nat. Rev. Neurosci. 2012, 13, 465–477. [Google Scholar] [CrossRef] [PubMed]
  90. Miranda, A.F.; Boegman, R.J.; Beninger, R.J.; Jhamandas, K. Protection against quinolinic acid-mediated excitotoxicity in nigrostriatal dopaminergic neurons by endogenous kynurenic acid. Neuroscience 1997, 78, 967–975. [Google Scholar] [CrossRef] [PubMed]
  91. Harris, C.A.; Miranda, A.F.; Tanguay, J.J.; Boegman, R.J.; Beninger, R.J.; Jhamandas, K. Modulation of striatal quinolinate neurotoxicity by elevation of endogenous brain kynurenic acid. Br. J. Pharmacol. 1998, 124, 391–399. [Google Scholar] [CrossRef] [PubMed]
  92. Cervenka, I.; Agudelo, L.Z.; Ruas, J.L. Kynurenines: Tryptophan’s metabolites in exercise, inflammation, and mental health. Science 2017, 357, eaaf9794. [Google Scholar] [CrossRef] [PubMed]
  93. García-Lara, L.; Pérez-Severiano, F.; González-Esquivel, D.; Elizondo, G.; Segovia, J. Absence of aryl hydrocarbon receptors increases endogenous kynurenic acid levels and protects mouse brain against excitotoxic insult and oxidative stress. J. Neurosci. Res. 2015, 93, 1423–1433. [Google Scholar] [CrossRef] [PubMed]
  94. Ferreira, F.S.; Biasibetti-Brendler, H.; Pierozan, P.; Schmitz, F.; Bertó, C.G.; Prezzi, C.A.; Manfredini, V.; Wyse, A.T.S. Kynurenic Acid Restores Nrf2 Levels and Prevents Quinolinic Acid-Induced Toxicity in Rat Striatal Slices. Mol. Neurobiol. 2018, 55, 8538–8549. [Google Scholar] [CrossRef] [PubMed]
  95. Ferreira, F.S.; Schmitz, F.; Marques, E.P.; Siebert, C.; Wyse, A.T.S. Intrastriatal Quinolinic Acid Administration Impairs Redox Homeostasis and Induces Inflammatory Changes: Prevention by Kynurenic Acid. Neurotox. Res. 2020, 38, 50–58. [Google Scholar] [CrossRef]
  96. Fujigaki, H.; Yamamoto, Y.; Saito, K. L-Tryptophan-kynurenine pathway enzymes are therapeutic target for neuropsychiatric diseases: Focus on cell type differences. Neuropharmacology 2017, 112, 264–274. [Google Scholar] [CrossRef]
  97. Lima, S.; Kumar, S.; Gawandi, V.; Momany, C.; Phillips, R.S. Crystal structure of the Homo sapiens kynureninase-3-hydroxyhippuric acid inhibitor complex: Insights into the molecular basis of kynureninase substrate specificity. J. Med. Chem. 2009, 52, 389–396. [Google Scholar] [CrossRef]
  98. Wilson, K.; Mole, D.J.; Binnie, M.; Homer, N.Z.; Zheng, X.; Yard, B.A.; Iredale, J.P.; Auer, M.; Webster, S.P. Bacterial expression of human kynurenine 3-monooxygenase: Solubility, activity, purification. Protein Expr. Purif. 2014, 95, 96–103. [Google Scholar] [CrossRef]
  99. Urbańska, E.M.; Chmiel-Perzyńska, I.; Perzyński, A.; Derkacz, M.; Owe-Larsson, B. Endogenous Kynurenic Acid and Neurotoxicity. In Handbook of Neurotoxicity; Kostrzewa, R.M., Ed.; Springer: Berlin/Heidelberg, Germany, 2021; pp. 1–31. [Google Scholar]
  100. Zinger, A.; Barcia, C.; Herrero, M.T.; Guillemin, G.J. The involvement of neuroinflammation and kynurenine pathway in Parkinson’s disease. Parkinsons Dis. 2011, 2011, 716859. [Google Scholar] [CrossRef]
  101. Boros, F.A.; Bohár, Z.; Vécsei, L. Genetic alterations affecting the genes encoding the enzymes of the kynurenine pathway and their association with human diseases. Mutat. Res. Rev. Mutat. Res. 2018, 776, 32–45. [Google Scholar] [CrossRef]
  102. Wirthgen, E.; Hoeflich, A.; Rebl, A.; Günther, J. Kynurenic Acid: The Janus-Faced Role of an Immunomodulatory Tryptophan Metabolite and Its Link to Pathological Conditions. Front. Immunol. 2017, 8, 1957. [Google Scholar] [CrossRef]
  103. Kacser, H.; Burns, J.A. The control of flux. In Rate Control of Biological Processes. Symposium of the Society for Experimental Biology; Davies, D.D., Ed.; Cambridge University Press: Cambridge, UK, 1973; Volume 27, pp. 65–104. [Google Scholar]
  104. Heinrich, R.; Rapoport, T.A. A linear steady-state treatment of enzymatic chains. General properties, control and effector strength. Eur. J. Biochem. 1974, 42, 89–95. [Google Scholar] [PubMed]
  105. Kell, D.B.; Westerhoff, H.V. Metabolic control theory: Its role in microbiology and biotechnology. FEMS Microbiol. Rev. 1986, 39, 305–320. [Google Scholar]
  106. Fell, D.A. Understanding the Control of Metabolism; Portland Press: London, UK, 1996. [Google Scholar]
  107. Fell, D.A.; Saavedra, E.; Rohwer, J. 50 years of Metabolic Control Analysis: Its past and current influence in the biological sciences. Biosystems 2024, 235, 105086. [Google Scholar] [CrossRef]
  108. Rios-Avila, L.; Nijhout, H.F.; Reed, M.C.; Sitren, H.S.; Gregory, J.F., 3rd. A mathematical model of tryptophan metabolism via the kynurenine pathway provides insights into the effects of vitamin B-6 deficiency, tryptophan loading, and induction of tryptophan 2,3-dioxygenase on tryptophan metabolites. J. Nutr. 2013, 143, 1509–1519. [Google Scholar] [CrossRef]
  109. Han, Q.; Cai, T.; Tagle, D.A.; Li, J. Structure, expression, and function of kynurenine aminotransferases in human and rodent brains. Cell Mol. Life Sci. 2010, 67, 353–368. [Google Scholar] [CrossRef]
  110. Bergmann, F.T.; Hoops, S.; Klahn, B.; Kummer, U.; Mendes, P.; Pahle, J.; Sahle, S. COPASI and its applications in biotechnology. J. Biotechnol. 2017, 261, 215–220. [Google Scholar] [CrossRef]
  111. Hoops, S.; Sahle, S.; Gauges, R.; Lee, C.; Pahle, J.; Simus, N.; Singhal, M.; Xu, L.; Mendes, P.; Kummer, U. COPASI: A COmplex PAthway SImulator. Bioinformatics 2006, 22, 3067–3074. [Google Scholar] [PubMed]
  112. Mendes, P.; Hoops, S.; Sahle, S.; Gauges, R.; Dada, J.; Kummer, U. Computational modeling of biochemical networks using COPASI. Methods Mol. Biol. 2009, 500, 17–59. [Google Scholar]
  113. Keating, S.M.; Waltemath, D.; Konig, M.; Zhang, F.; Drager, A.; Chaouiya, C.; Bergmann, F.T.; Finney, A.; Gillespie, C.S.; Helikar, T.; et al. SBML Level 3: An extensible format for the exchange and reuse of biological models. Mol. Syst. Biol. 2020, 16, e9110. [Google Scholar] [CrossRef]
  114. Thiele, I.; Swainston, N.; Fleming, R.M.T.; Hoppe, A.; Sahoo, S.; Aurich, M.K.; Haraldsdottír, H.; Mo, M.L.; Rolfsson, O.; Stobbe, M.D.; et al. A community-driven global reconstruction of human metabolism. Nat. Biotechnol. 2013, 31, 419–425. [Google Scholar] [CrossRef] [PubMed]
  115. Smallbone, K. Striking a balance with Recon 2.1. arXiv 2013, arXiv:1311.5696. [Google Scholar]
  116. Herrgård, M.J.; Swainston, N.; Dobson, P.; Dunn, W.B.; Arga, K.Y.; Arvas, M.; Blüthgen, N.; Borger, S.; Costenoble, R.; Heinemann, M.; et al. A consensus yeast metabolic network obtained from a community approach to systems biology. Nat. Biotechnol. 2008, 26, 1155–1160. [Google Scholar] [PubMed]
  117. Ball, H.J.; Jusof, F.F.; Bakmiwewa, S.M.; Hunt, N.H.; Yuasa, H.J. Tryptophan-catabolizing enzymes-party of three. Front. Immunol. 2014, 5, 485. [Google Scholar] [CrossRef]
  118. Yuasa, H.J.; Ushigoe, A.; Ball, H.J. Molecular evolution of bacterial indoleamine 2,3-dioxygenase. Gene 2011, 485, 22–31. [Google Scholar] [CrossRef] [PubMed]
  119. Yuasa, H.J.; Ball, H.J. Molecular evolution and characterization of fungal indoleamine 2,3-dioxygenases. J. Mol. Evol. 2011, 72, 160–168. [Google Scholar] [CrossRef] [PubMed]
  120. Yuasa, H.J.; Ball, H.J. The evolution of three types of indoleamine 2,3 dioxygenases in fungi with distinct molecular and biochemical characteristics. Gene 2012, 504, 64–74. [Google Scholar] [CrossRef] [PubMed]
  121. Yuasa, H.J.; Ball, H.J. Indoleamine 2,3-dioxygenases with very low catalytic activity are well conserved across kingdoms: IDOs of Basidiomycota. Fungal Genet. Biol. 2013, 56, 98–106. [Google Scholar] [CrossRef] [PubMed]
  122. Yuasa, H.J.; Ball, H.J. Efficient tryptophan-catabolizing activity is consistently conserved through evolution of TDO enzymes, but not IDO enzymes. J. Exp. Zool. B Mol. Dev. Evol. 2015, 324, 128–140. [Google Scholar] [CrossRef] [PubMed]
  123. Yuasa, H.J.; Sugiura, M.; Harumoto, T. A single amino acid residue regulates the substrate affinity and specificity of indoleamine 2,3-dioxygenase. Arch. Biochem. Biophys. 2018, 640, 1–9. [Google Scholar] [CrossRef] [PubMed]
  124. Lima, W.C.; Varani, A.M.; Menck, C.F. NAD biosynthesis evolution in bacteria: Lateral gene transfer of kynurenine pathway in Xanthomonadales and Flavobacteriales. Mol. Biol. Evol. 2009, 26, 399–406. [Google Scholar] [CrossRef] [PubMed]
  125. Ball, H.J.; Sanchez-Perez, A.; Weiser, S.; Austin, C.J.; Astelbauer, F.; Miu, J.; McQuillan, J.A.; Stocker, R.; Jermiin, L.S.; Hunt, N.H. Characterization of an indoleamine 2,3-dioxygenase-like protein found in humans and mice. Gene 2007, 396, 203–213. [Google Scholar] [CrossRef] [PubMed]
  126. Metz, R.; Duhadaway, J.B.; Kamasani, U.; Laury-Kleintop, L.; Muller, A.J.; Prendergast, G.C. Novel tryptophan catabolic enzyme IDO2 is the preferred biochemical target of the antitumor indoleamine 2,3-dioxygenase inhibitory compound D-1-methyl-tryptophan. Cancer Res. 2007, 67, 7082–7087. [Google Scholar] [CrossRef] [PubMed]
  127. Theate, I.; van Baren, N.; Pilotte, L.; Moulin, P.; Larrieu, P.; Renauld, J.C.; Herve, C.; Gutierrez-Roelens, I.; Marbaix, E.; Sempoux, C.; et al. Extensive profiling of the expression of the indoleamine 2,3-dioxygenase 1 protein in normal and tumoral human tissues. Cancer Immunol. Res. 2015, 3, 161–172. [Google Scholar] [CrossRef]
  128. Yamazaki, F.; Kuroiwa, T.; Takikawa, O.; Kido, R. Human indolylamine 2,3-dioxygenase. Its tissue distribution, and characterization of the placental enzyme. Biochem. J. 1985, 230, 635–638. [Google Scholar] [CrossRef] [PubMed]
  129. Löb, S.; Konigsrainer, A.; Zieker, D.; Brücher, B.L.; Rammensee, H.G.; Opelz, G.; Terness, P. IDO1 and IDO2 are expressed in human tumors: Levo- but not dextro-1-methyl tryptophan inhibits tryptophan catabolism. Cancer Immunol. Immunother. 2009, 58, 153–157. [Google Scholar] [CrossRef]
  130. Girithar, H.N.; Staats Pires, A.; Ahn, S.B.; Guillemin, G.J.; Gluch, L.; Heng, B. Involvement of the kynurenine pathway in breast cancer: Updates on clinical research and trials. Br. J. Cancer 2023, 129, 185–203. [Google Scholar] [CrossRef]
  131. Munn, D.H.; Zhou, M.; Attwood, J.T.; Bondarev, I.; Conway, S.J.; Marshall, B.; Brown, C.; Mellor, A.L. Prevention of allogeneic fetal rejection by tryptophan catabolism. Science 1998, 281, 1191–1193. [Google Scholar] [CrossRef]
  132. Croitoru-Lamoury, J.; Lamoury, F.M.; Caristo, M.; Suzuki, K.; Walker, D.; Takikawa, O.; Taylor, R.; Brew, B.J. Interferon-gamma regulates the proliferation and differentiation of mesenchymal stem cells via activation of indoleamine 2,3 dioxygenase (IDO). PLoS ONE 2011, 6, e14698. [Google Scholar] [CrossRef]
  133. Sarkar, S.A.; Wong, R.; Hackl, S.I.; Moua, O.; Gill, R.G.; Wiseman, A.; Davidson, H.W.; Hutton, J.C. Induction of indoleamine 2,3-dioxygenase by interferon-gamma in human islets. Diabetes 2007, 56, 72–79. [Google Scholar] [CrossRef] [PubMed]
  134. Ganesan, S.; Roy, C.R. Host cell depletion of tryptophan by IFNgamma-induced Indoleamine 2,3-dioxygenase 1 (IDO1) inhibits lysosomal replication of Coxiella burnetii. PLoS Pathog. 2019, 15, e1007955. [Google Scholar] [CrossRef]
  135. Capece, L.; Arrar, M.; Roitberg, A.E.; Yeh, S.R.; Marti, M.A.; Estrin, D.A. Substrate stereo-specificity in tryptophan dioxygenase and indoleamine 2,3-dioxygenase. Proteins 2010, 78, 2961–2972. [Google Scholar] [CrossRef]
  136. Shimizu, T.; Nomiyama, S.; Hirata, F.; Hayaishi, O. Indoleamine 2,3-dioxygenase. Purification and some properties. J. Biol. Chem. 1978, 253, 4700–4706. [Google Scholar]
  137. Geng, J.; Liu, A. Heme-dependent dioxygenases in tryptophan oxidation. Arch. Biochem. Biophys. 2014, 544, 18–26. [Google Scholar] [CrossRef]
  138. Lewis-Ballester, A.; Forouhar, F.; Kim, S.M.; Lew, S.; Wang, Y.; Karkashon, S.; Seetharaman, J.; Batabyal, D.; Chiang, B.Y.; Hussain, M.; et al. Molecular basis for catalysis and substrate-mediated cellular stabilization of human tryptophan 2,3-dioxygenase. Sci. Rep. 2016, 6, 35169. [Google Scholar] [CrossRef]
  139. Pantouris, G.; Serys, M.; Yuasa, H.J.; Ball, H.J.; Mowat, C.G. Human indoleamine 2,3-dioxygenase-2 has substrate specificity and inhibition characteristics distinct from those of indoleamine 2,3-dioxygenase-1. Amino Acids 2014, 46, 2155–2163. [Google Scholar] [CrossRef] [PubMed]
  140. Geisler, S.; Mayersbach, P.; Becker, K.; Schennach, H.; Fuchs, D.; Gostner, J.M. Serum tryptophan, kynurenine, phenylalanine, tyrosine and neopterin concentrations in 100 healthy blood donors. Pteridines 2015, 26, 31–36. [Google Scholar] [CrossRef]
  141. Perez-Castro, L.; Garcia, R.; Venkateswaran, N.; Barnes, S.; Conacci-Sorrell, M. Tryptophan and its metabolites in normal physiology and cancer etiology. FEBS J. 2023, 290, 7–27. [Google Scholar] [CrossRef]
  142. Kudo, Y.; Koh, I.; Sugimoto, J. Localization of Indoleamine 2,3-Dioxygenase-1 and Indoleamine 2,3-Dioxygenase-2 at the Human Maternal-Fetal Interface. Int. J. Tryptophan Res. 2020, 13, 1178646920984163. [Google Scholar] [CrossRef] [PubMed]
  143. Mandarano, M.; Bellezza, G.; Belladonna, M.L.; Vannucci, J.; Gili, A.; Ferri, I.; Lupi, C.; Ludovini, V.; Falabella, G.; Metro, G.; et al. Indoleamine 2,3-Dioxygenase 2 Immunohistochemical Expression in Resected Human Non-small Cell Lung Cancer: A Potential New Prognostic Tool. Front. Immunol. 2020, 11, 839. [Google Scholar] [CrossRef]
  144. Trabanelli, S.; Ocadlikova, D.; Ciciarello, M.; Salvestrini, V.; Lecciso, M.; Jandus, C.; Metz, R.; Evangelisti, C.; Laury-Kleintop, L.; Romero, P.; et al. The SOCS3-independent expression of IDO2 supports the homeostatic generation of T regulatory cells by human dendritic cells. J. Immunol. 2014, 192, 1231–1240. [Google Scholar] [CrossRef]
  145. Merlo, L.M.F.; DuHadaway, J.B.; Montgomery, J.D.; Peng, W.D.; Murray, P.J.; Prendergast, G.C.; Caton, A.J.; Muller, A.J.; Mandik-Nayak, L. Differential Roles of IDO1 and IDO2 in T and B Cell Inflammatory Immune Responses. Front. Immunol. 2020, 11, 1861. [Google Scholar] [CrossRef]
  146. Meng, B.; Wu, D.; Gu, J.; Ouyang, S.; Ding, W.; Liu, Z.J. Structural and functional analyses of human tryptophan 2,3-dioxygenase. Proteins 2014, 82, 3210–3216. [Google Scholar] [CrossRef]
  147. Mellor, A.L.; Munn, D.H. IDO expression by dendritic cells: Tolerance and tryptophan catabolism. Nat. Rev. Immunol. 2004, 4, 762–774. [Google Scholar] [CrossRef] [PubMed]
  148. O’Hagan, S.; Wright Muelas, M.; Day, P.J.; Lundberg, E.; Kell, D.B. GeneGini: Assessment via the Gini coefficient of reference ‘‘housekeeping’’ genes and diverse human transporter expression profiles. Cell Syst. 2018, 6, 230–244. [Google Scholar]
  149. Klaessens, S.; Stroobant, V.; De Plaen, E.; Van den Eynde, B.J. Systemic tryptophan homeostasis. Front. Mol. Biosci. 2022, 9, 897929. [Google Scholar] [CrossRef]
  150. Van Baren, N.; Van den Eynde, B.J. Tryptophan-degrading enzymes in tumoral immune resistance. Front. Immunol. 2015, 6, 34. [Google Scholar] [CrossRef]
  151. Hoffmann, D.; Dvorakova, T.; Stroobant, V.; Bouzin, C.; Daumerie, A.; Solvay, M.; Klaessens, S.; Letellier, M.C.; Renauld, J.C.; van Baren, N.; et al. Tryptophan 2,3-Dioxygenase Expression Identified in Human Hepatocellular Carcinoma Cells and in Intratumoral Pericytes of Most Cancers. Cancer Immunol. Res. 2020, 8, 19–31. [Google Scholar] [CrossRef]
  152. Boros, F.A.; Vécsei, L. Tryptophan 2,3-dioxygenase, a novel therapeutic target for Parkinson’s disease. Expert. Opin. Ther. Targets 2021, 25, 877–888. [Google Scholar] [CrossRef]
  153. Fathi, M.; Vakili, K.; Yaghoobpoor, S.; Tavasol, A.; Jazi, K.; Hajibeygi, R.; Shool, S.; Sodeifian, F.; Klegeris, A.; McElhinney, A.; et al. Dynamic changes in metabolites of the kynurenine pathway in Alzheimer’s disease, Parkinson’s disease, and Huntington’s disease: A systematic Review and meta-analysis. Front. Immunol. 2022, 13, 997240. [Google Scholar] [CrossRef]
  154. Basran, J.; Rafice, S.A.; Chauhan, N.; Efimov, I.; Cheesman, M.R.; Ghamsari, L.; Raven, E.L. A kinetic, spectroscopic, and redox study of human tryptophan 2,3-dioxygenase. Biochemistry 2008, 47, 4752–4760. [Google Scholar] [CrossRef] [PubMed]
  155. Pabarcus, M.K.; Casida, J.E. Cloning, expression, and catalytic triad of recombinant arylformamidase. Protein Expr. Purif. 2005, 44, 39–44. [Google Scholar] [CrossRef]
  156. Madeira, F.; Pearce, M.; Tivey, A.R.N.; Basutkar, P.; Lee, J.; Edbali, O.; Madhusoodanan, N.; Kolesnikov, A.; Lopez, R. Search and sequence analysis tools services from EMBL-EBI in 2022. Nucleic Acids Res. 2022, 50, W276–W279. [Google Scholar] [CrossRef]
  157. Schuettengruber, B.; Doetzlhofer, A.; Kroboth, K.; Wintersberger, E.; Seiser, C. Alternate activation of two divergently transcribed mouse genes from a bidirectional promoter is linked to changes in histone modification. J. Biol. Chem. 2003, 278, 1784–1793. [Google Scholar] [CrossRef] [PubMed]
  158. Dobrovolsky, V.N.; Bowyer, J.F.; Pabarcus, M.K.; Heflich, R.H.; Williams, L.D.; Doerge, D.R.; Arvidsson, B.; Bergquist, J.; Casida, J.E. Effect of arylformamidase (kynurenine formamidase) gene inactivation in mice on enzymatic activity, kynurenine pathway metabolites and phenotype. Biochim. Biophys. Acta 2005, 1724, 163–172. [Google Scholar] [CrossRef] [PubMed]
  159. Wogulis, M.; Chew, E.R.; Donohoue, P.D.; Wilson, D.K. Identification of formyl kynurenine formamidase and kynurenine aminotransferase from Saccharomyces cerevisiae using crystallographic, bioinformatic and biochemical evidence. Biochemistry 2008, 47, 1608–1621. [Google Scholar] [CrossRef]
  160. Moscioni, A.D.; Engel, J.L.; Casida, J.E. Kynurenine formamidase inhibition as a possible mechanism for certain teratogenic effects of organophosphorus and methylcarbamate insecticides in chicken embryos. Biochem. Pharmacol. 1977, 26, 2251–2258. [Google Scholar] [CrossRef] [PubMed]
  161. Han, Q.; Robinson, H.; Li, J. Biochemical identification and crystal structure of kynurenine formamidase from Drosophila melanogaster. Biochem. J. 2012, 446, 253–260. [Google Scholar] [CrossRef]
  162. Seifert, J.; Pewnim, T. Alteration of mice L-tryptophan metabolism by the organophosphorous acid triester diazinon. Biochem. Pharmacol. 1992, 44, 2243–2250. [Google Scholar] [CrossRef]
  163. Casida, J.E.; Quistad, G.B. Serine hydrolase targets of organophosphorus toxicants. Chem. Biol. Interact. 2005, 157–158, 277–283. [Google Scholar] [CrossRef]
  164. Schwarcz, R.; Pellicciari, R. Manipulation of Brain Kynurenines: Glial Targets, Neuronal Effects, and Clinical Opportunities. J. Pharm. Exp. Therapeut 2002, 303, 1010. [Google Scholar]
  165. Bellocchi, D.; Macchiarulo, A.; Carotti, A.; Pellicciari, R. Quantum mechanics/molecular mechanics (QM/MM) modeling of the irreversible transamination of L-kynurenine to kynurenic acid: The round dance of kynurenine aminotransferase II. Biochim. Biophys. Acta 2009, 1794, 1802–1812. [Google Scholar] [CrossRef]
  166. Koper, K.; Han, S.W.; Pastor, D.C.; Yoshikuni, Y.; Maeda, H.A. Evolutionary origin and functional diversification of aminotransferases. J. Biol. Chem. 2022, 298, 102122. [Google Scholar] [CrossRef]
  167. Bulfer, S.L.; Brunzelle, J.S.; Trievel, R.C. Crystal structure of Saccharomyces cerevisiae Aro8, a putative alpha-aminoadipate aminotransferase. Protein Sci. 2013, 22, 1417–1424. [Google Scholar] [CrossRef]
  168. Han, Q.; Fang, J.; Li, J. Kynurenine aminotransferase and glutamine transaminase K of Escherichia coli: Identity with aspartate aminotransferase. Biochem. J. 2001, 360, 617–623. [Google Scholar] [CrossRef]
  169. Ohashi, K.; Chaleckis, R.; Takaine, M.; Wheelock, C.E.; Yoshida, S. Kynurenine aminotransferase activity of Aro8/Aro9 engage tryptophan degradation by producing kynurenic acid in Saccharomyces cerevisiae. Sci. Rep. 2017, 7, 12180. [Google Scholar] [CrossRef]
  170. Radi, M.S.; Sora, J.E.S.; Kim, S.H.; Sudarsan, S.; Sastry, A.V.; Kell, D.B.; Herrgård, M.J.; Feist, A.M. Membrane transporter identification and modulation via adaptive laboratory evolution. Metab. Eng. 2022, 72, 376–390. [Google Scholar]
  171. Genestet, C.; Le Gouellec, A.; Chaker, H.; Polack, B.; Guery, B.; Toussaint, B.; Stasia, M.J. Scavenging of reactive oxygen species by tryptophan metabolites helps Pseudomonas aeruginosa escape neutrophil killing. Free Radic. Biol. Med. 2014, 73, 400–410. [Google Scholar] [CrossRef]
  172. Bortolotti, P.; Hennart, B.; Thieffry, C.; Jausions, G.; Faure, E.; Grandjean, T.; Thepaut, M.; Dessein, R.; Allorge, D.; Guery, B.P.; et al. Tryptophan catabolism in Pseudomonas aeruginosa and potential for inter-kingdom relationship. BMC Microbiol. 2016, 16, 137. [Google Scholar] [CrossRef]
  173. Jansen, R.S.; Mandyoli, L.; Hughes, R.; Wakabayashi, S.; Pinkham, J.T.; Selbach, B.; Guinn, K.M.; Rubin, E.J.; Sacchettini, J.C.; Rhee, K.Y. Aspartate aminotransferase Rv3722c governs aspartate-dependent nitrogen metabolism in Mycobacterium tuberculosis. Nat. Commun. 2020, 11, 1960. [Google Scholar] [CrossRef]
  174. Han, Q.; Robinson, H.; Cai, T.; Tagle, D.A.; Li, J. Structural insight into the inhibition of human kynurenine aminotransferase I/glutamine transaminase K. J. Med. Chem. 2009, 52, 2786–2793. [Google Scholar] [CrossRef]
  175. Nadvi, N.A.; Salam, N.K.; Park, J.; Akladios, F.N.; Kapoor, V.; Collyer, C.A.; Gorrell, M.D.; Church, W.B. High resolution crystal structures of human kynurenine aminotransferase-I bound to PLP cofactor, and in complex with aminooxyacetate. Protein Sci. 2017, 26, 727–736. [Google Scholar] [CrossRef] [PubMed]
  176. Rossi, F.; Garavaglia, S.; Montalbano, V.; Walsh, M.A.; Rizzi, M. Crystal structure of human kynurenine aminotransferase II, a drug target for the treatment of schizophrenia. J. Biol. Chem. 2008, 283, 3559–3566. [Google Scholar] [CrossRef] [PubMed]
  177. Han, Q.; Cai, T.; Tagle, D.A.; Robinson, H.; Li, J. Substrate specificity and structure of human aminoadipate aminotransferase/kynurenine aminotransferase II. Biosci. Rep. 2008, 28, 205–215. [Google Scholar] [CrossRef]
  178. Tuttle, J.B.; Anderson, M.; Bechle, B.M.; Campbell, B.M.; Chang, C.; Dounay, A.B.; Evrard, E.; Fonseca, K.R.; Gan, X.; Ghosh, S.; et al. Structure-Based Design of Irreversible Human KAT II Inhibitors: Discovery of New Potency-Enhancing Interactions. ACS Med. Chem. Lett. 2013, 4, 37–40. [Google Scholar] [CrossRef]
  179. Nematollahi, A.; Sun, G.; Harrop, S.J.; Hanrahan, J.R.; Church, W.B. Structure of the PLP-Form of the Human Kynurenine Aminotransferase II in a Novel Spacegroup at 1.83 Å Resolution. Int. J. Mol. Sci. 2016, 17, 446. [Google Scholar] [CrossRef]
  180. Nematollahi, A.; Church, W.B.; Nadvi, N.A.; Gorrell, M.D.; Sun, G. Homology modeling of human kynurenine aminotransferase III and observations on inhibitor binding using molecular docking. Cent. Nerv. Syst. Agents Med. Chem. 2014, 14, 2–9. [Google Scholar] [CrossRef] [PubMed]
  181. Jiang, X.; Wang, J.; Chang, H.; Zhou, Y. Recombinant expression, purification and crystallographic studies of the mature form of human mitochondrial aspartate aminotransferase. Biosci. Trends 2016, 10, 79–84. [Google Scholar] [CrossRef] [PubMed]
  182. Cooper, A.J.; Pinto, J.T.; Krasnikov, B.F.; Niatsetskaya, Z.V.; Han, Q.; Li, J.; Vauzour, D.; Spencer, J.P. Substrate specificity of human glutamine transaminase K as an aminotransferase and as a cysteine S-conjugate beta-lyase. Arch. Biochem. Biophys. 2008, 474, 72–81. [Google Scholar] [CrossRef]
  183. Badillo-Ramírez, I.; Saniger, J.M.; Rivas-Arancibia, S. 5-S-cysteinyl-dopamine, a neurotoxic endogenous metabolite of dopamine: Implications for Parkinson’s disease. Neurochem. Int. 2019, 129, 104514. [Google Scholar] [CrossRef]
  184. Uhlén, M.; Fagerberg, L.; Hallstrom, B.M.; Lindskog, C.; Oksvold, P.; Mardinoglu, A.; Sivertsson, Ǻ.; Kampf, C.; Sjöstedt, E.; Asplund, A.; et al. Tissue-based map of the human proteome. Science 2015, 347, 1260419. [Google Scholar] [CrossRef]
  185. Sjöstedt, E.; Zhong, W.; Fagerberg, L.; Karlsson, M.; Mitsios, N.; Adori, C.; Oksvold, P.; Edfors, F.; Limiszewska, A.; Hikmet, F.; et al. An atlas of the protein-coding genes in the human, pig, and mouse brain. Science 2020, 367, eaay5947. [Google Scholar] [CrossRef]
  186. Han, Q.; Li, J.; Li, J. pH dependence, substrate specificity and inhibition of human kynurenine aminotransferase I. Eur. J. Biochem. 2004, 271, 4804–4814. [Google Scholar] [CrossRef] [PubMed]
  187. Cooper, A.J.; Shurubor, Y.I.; Dorai, T.; Pinto, J.T.; Isakova, E.P.; Deryabina, Y.I.; Denton, T.T.; Krasnikov, B.F. omega-Amidase: An underappreciated, but important enzyme in L-glutamine and L-asparagine metabolism; relevance to sulfur and nitrogen metabolism, tumor biology and hyperammonemic diseases. Amino Acids 2016, 48, 1–20. [Google Scholar] [CrossRef] [PubMed]
  188. Potter, M.C.; Elmer, G.I.; Bergeron, R.; Albuquerque, E.X.; Guidetti, P.; Wu, H.Q.; Schwarcz, R. Reduction of endogenous kynurenic acid formation enhances extracellular glutamate, hippocampal plasticity, and cognitive behavior. Neuropsychopharmacology 2010, 35, 1734–1742. [Google Scholar] [CrossRef]
  189. Kapoor, R.; Lim, K.S.; Cheng, A.; Garrick, T.; Kapoor, V. Preliminary evidence for a link between schizophrenia and NMDA-glycine site receptor ligand metabolic enzymes, d-amino acid oxidase (DAAO) and kynurenine aminotransferase-1 (KAT-1). Brain Res. 2006, 1106, 205–210. [Google Scholar] [CrossRef]
  190. Barazorda-Ccahuana, H.L.; Zevallos-Delgado, C.; Valencia, D.E.; Gómez, B. Molecular Dynamics Simulation of Kynurenine AminotransferaseType II with Nicotine as a Ligand: A Possible Biochemical Role ofNicotine in Schizophrenia. ACS Omega 2019, 4, 710–717. [Google Scholar] [CrossRef]
  191. Chang, C.; Fonseca, K.R.; Li, C.; Horner, W.; Zawadzke, L.E.; Salafia, M.A.; Welch, K.A.; Strick, C.A.; Campbell, B.M.; Gernhardt, S.S.; et al. Quantitative Translational Analysis of Brain Kynurenic Acid Modulation via Irreversible Kynurenine Aminotransferase II Inhibition. Mol. Pharmacol. 2018, 94, 823–833. [Google Scholar] [CrossRef]
  192. Dounay, A.B.; Anderson, M.; Bechle, B.M.; Campbell, B.M.; Claffey, M.M.; Evdokimov, A.; Evrard, E.; Fonseca, K.R.; Gan, X.; Ghosh, S.; et al. Discovery of Brain-Penetrant, Irreversible Kynurenine Aminotransferase II Inhibitors for Schizophrenia. ACS Med. Chem. Lett. 2012, 3, 187–192. [Google Scholar] [CrossRef]
  193. Dounay, A.B.; Anderson, M.; Bechle, B.M.; Evrard, E.; Gan, X.; Kim, J.Y.; McAllister, L.A.; Pandit, J.; Rong, S.; Salafia, M.A.; et al. PF-04859989 as a template for structure-based drug design: Identification of new pyrazole series of irreversible KAT II inhibitors with improved lipophilic efficiency. Bioorg Med. Chem. Lett. 2013, 23, 1961–1966. [Google Scholar] [CrossRef] [PubMed]
  194. Nematollahi, A.; Sun, G.; Jayawickrama, G.S.; Church, W.B. Kynurenine Aminotransferase Isozyme Inhibitors: A Review. Int. J. Mol. Sci. 2016, 17, 946. [Google Scholar] [CrossRef]
  195. Yoshida, Y.; Fujigaki, H.; Kato, K.; Yamazaki, K.; Fujigaki, S.; Kunisawa, K.; Yamamoto, Y.; Mouri, A.; Oda, A.; Nabeshima, T.; et al. Selective and competitive inhibition of kynurenine aminotransferase 2 by glycyrrhizic acid and its analogues. Sci. Rep. 2019, 9, 10243. [Google Scholar] [CrossRef]
  196. Erhardt, S.; Schwieler, L.; Nilsson, L.; Linderholm, K.; Engberg, G. The kynurenic acid hypothesis of schizophrenia. Physiol. Behav. 2007, 92, 203–209. [Google Scholar] [CrossRef] [PubMed]
  197. Birner, A.; Platzer, M.; Bengesser, S.A.; Dalkner, N.; Fellendorf, F.T.; Queissner, R.; Pilz, R.; Rauch, P.; Maget, A.; Hamm, C.; et al. Increased breakdown of kynurenine towards its neurotoxic branch in bipolar disorder. PLoS ONE 2017, 12, e0172699. [Google Scholar] [CrossRef]
  198. Bai, M.Y.; Lovejoy, D.B.; Guillemin, G.J.; Kozak, R.; Stone, T.W.; Koola, M.M. Galantamine-Memantine Combination and Kynurenine Pathway Enzyme Inhibitors in the Treatment of Neuropsychiatric Disorders. Complex. Psychiatry 2021, 7, 19–33. [Google Scholar] [CrossRef] [PubMed]
  199. Fukuwatari, T. Possibility of Amino Acid Treatment to Prevent the Psychiatric Disorders via Modulation of the Production of Tryptophan Metabolite Kynurenic Acid. Nutrients 2020, 12, 1403. [Google Scholar] [CrossRef]
  200. Yu, P.; Di Prospero, N.A.; Sapko, M.T.; Cai, T.; Chen, A.; Melendez-Ferro, M.; Du, F.; Whetsell, W.O., Jr.; Guidetti, P.; Schwarcz, R.; et al. Biochemical and phenotypic abnormalities in kynurenine aminotransferase II-deficient mice. Mol. Cell Biol. 2004, 24, 6919–6930. [Google Scholar] [CrossRef]
  201. Hallen, A.; Jamie, J.F.; Cooper, A.J. Lysine metabolism in mammalian brain: An update on the importance of recent discoveries. Amino Acids 2013, 45, 1249–1272. [Google Scholar] [CrossRef]
  202. Madadi, S.; Pasbakhsh, P.; Tahmasebi, F.; Mortezaee, K.; Khanehzad, M.; Boroujeni, F.B.; Noorzehi, G.; Kashani, I.R. Astrocyte ablation induced by La-aminoadipate (L-AAA) potentiates remyelination in a cuprizone demyelinating mouse model. Metab. Brain Dis. 2019, 34, 593–603. [Google Scholar] [CrossRef]
  203. Okuno, E.; Tsujimoto, M.; Nakamura, M.; Kido, R. 2-Aminoadipate-2-oxoglutarate aminotransferase isoenzymes in human liver: A plausible physiological role in lysine and tryptophan metabolism. Enzyme Protein 1993, 47, 136–148. [Google Scholar] [CrossRef]
  204. Tanaka, M.; Szabó, Á.; Spekker, E.; Polyák, H.; Tóth, F.; Vécsei, L. Mitochondrial Impairment: A Common Motif in Neuropsychiatric Presentation? The Link to the Tryptophan-Kynurenine Metabolic System. Cells 2022, 11, 2607. [Google Scholar] [CrossRef]
  205. Yang, C.; Zhang, L.; Han, Q.; Liao, C.; Lan, J.; Ding, H.; Zhou, H.; Diao, X.; Li, J. Kynurenine aminotransferase 3/glutamine transaminase L/cysteine conjugate beta-lyase 2 is a major glutamine transaminase in the mouse kidney. Biochem. Biophys. Rep. 2016, 8, 234–241. [Google Scholar] [CrossRef]
  206. Wu, Q.; Huang, J.; Wu, R. Drugs Based on NMDAR Hypofunction Hypothesis in Schizophrenia. Front. Neurosci. 2021, 15, 641047. [Google Scholar] [CrossRef]
  207. Jayawickrama, G.S.; Sadig, R.R.; Sun, G.; Nematollahi, A.; Nadvi, N.A.; Hanrahan, J.R.; Gorrell, M.D.; Church, W.B. Kynurenine Aminotransferases and the Prospects of Inhibitors for the Treatment of Schizophrenia. Curr. Med. Chem. 2015, 22, 2902–2918. [Google Scholar] [CrossRef] [PubMed]
  208. Yu, P.; Li, Z.; Zhang, L.; Tagle, D.A.; Cai, T. Characterization of kynurenine aminotransferase III, a novel member of a phylogenetically conserved KAT family. Gene 2006, 365, 111–118. [Google Scholar] [CrossRef] [PubMed]
  209. Han, Q.; Robinson, H.; Cai, T.; Tagle, D.A.; Li, J. Biochemical and structural properties of mouse kynurenine aminotransferase III. Mol. Cell Biol. 2009, 29, 784–793. [Google Scholar] [CrossRef]
  210. Monné, M.; Miniero, D.V.; Iacobazzi, V.; Bisaccia, F.; Fiermonte, G. The mitochondrial oxoglutarate carrier: From identification to mechanism. J. Bioenerg. Biomembr. 2013, 45, 1–13. [Google Scholar] [CrossRef]
  211. Guidetti, P.; Amori, L.; Sapko, M.T.; Okuno, E.; Schwarcz, R. Mitochondrial aspartate aminotransferase: A third kynurenate-producing enzyme in the mammalian brain. J. Neurochem. 2007, 102, 103–111. [Google Scholar] [CrossRef]
  212. Han, Q.; Robinson, H.; Cai, T.; Tagle, D.A.; Li, J. Biochemical and structural characterization of mouse mitochondrial aspartate aminotransferase, a newly identified kynurenine aminotransferase-IV. Biosci. Rep. 2011, 31, 323–332. [Google Scholar] [CrossRef] [PubMed]
  213. Agudelo, L.Z.; Ferreira, D.M.S.; Cervenka, I.; Bryzgalova, G.; Dadvar, S.; Jannig, P.R.; Pettersson-Klein, A.T.; Lakshmikanth, T.; Sustarsic, E.G.; Porsmyr-Palmertz, M.; et al. Kynurenic Acid and Gpr35 Regulate Adipose Tissue Energy Homeostasis and Inflammation. Cell Metab. 2018, 27, 378–392.e375. [Google Scholar] [CrossRef] [PubMed]
  214. Palzkill, V.R.; Thome, T.; Murillo, A.L.; Khattri, R.B.; Ryan, T.E. Increasing plasma L-kynurenine impairs mitochondrial oxidative phosphorylation prior to the development of atrophy in murine skeletal muscle: A pilot study. Front. Physiol. 2022, 13, 992413. [Google Scholar] [CrossRef]
  215. Schlittler, M.; Goiny, M.; Agudelo, L.Z.; Venckunas, T.; Brazaitis, M.; Skurvydas, A.; Kamandulis, S.; Ruas, J.L.; Erhardt, S.; Westerblad, H.; et al. Endurance exercise increases skeletal muscle kynurenine aminotransferases and plasma kynurenic acid in humans. Am. J. Physiol. Cell Physiol. 2016, 310, C836–C840. [Google Scholar] [CrossRef]
  216. Alam, S.; Doherty, E.; Ortega-Prieto, P.; Arizanova, J.; Fets, L. Membrane transporters in cell physiology, cancer metabolism and drug response. Dis. Model. Mech. 2023, 16, dmm050404. [Google Scholar] [CrossRef] [PubMed]
  217. Kell, D.B.; Swainston, N.; Pir, P.; Oliver, S.G. Membrane transporter engineering in industrial biotechnology and whole-cell biocatalysis. Trends Biotechnol. 2015, 33, 237–246. [Google Scholar] [PubMed]
  218. Dobson, P.D.; Kell, D.B. Carrier-mediated cellular uptake of pharmaceutical drugs: An exception or the rule? Nat. Rev. Drug Disc. 2008, 7, 205–220. [Google Scholar]
  219. Kell, D.B.; Dobson, P.D. The cellular uptake of pharmaceutical drugs is mainly carrier-mediated and is thus an issue not so much of biophysics but of systems biology. In Proceedings of the Institut Beilstein Symposium on Systems Chemistry, Bozen, Italy, 26–30 May 2008; Hicks, M.G., Kettner, C., Eds.; Logos Verlag: Berlin, Germany, 2009; pp. 149–168. [Google Scholar]
  220. Kell, D.B.; Dobson, P.D.; Oliver, S.G. Pharmaceutical drug transport: The issues and the implications that it is essentially carrier-mediated only. Drug Disc. Today 2011, 16, 704–714. [Google Scholar]
  221. Kell, D.B.; Dobson, P.D.; Bilsland, E.; Oliver, S.G. The promiscuous binding of pharmaceutical drugs and their transporter-mediated uptake into cells: What we (need to) know and how we can do so. Drug Disc. Today 2013, 18, 218–239. [Google Scholar]
  222. Kell, D.B. Finding novel pharmaceuticals in the systems biology era using multiple effective drug targets, phenotypic screening, and knowledge of transporters: Where drug discovery went wrong and how to fix it. FEBS J. 2013, 280, 5957–5980. [Google Scholar] [PubMed]
  223. Kell, D.B.; Oliver, S.G. How drugs get into cells: Tested and testable predictions to help discriminate between transporter-mediated uptake and lipoidal bilayer diffusion. Front. Pharmacol. 2014, 5, 231. [Google Scholar]
  224. Kell, D.B. The transporter-mediated cellular uptake of pharmaceutical drugs is based on their metabolite-likeness and not on their bulk biophysical properties: Towards a systems pharmacology. Perspect. Sci. 2015, 6, 66–83. [Google Scholar] [CrossRef]
  225. Kell, D.B. How drugs pass through biological cell membranes—A paradigm shift in our understanding? Beilstein Mag. 2016, 2. [Google Scholar] [CrossRef]
  226. Giacomini, K.M.; Huang, S.M.; Tweedie, D.J.; Benet, L.Z.; Brouwer, K.L.; Chu, X.; Dahlin, A.; Evers, R.; Fischer, V.; Hillgren, K.M.; et al. Membrane transporters in drug development. Nat. Rev. Drug Discov. 2010, 9, 215–236. [Google Scholar] [PubMed]
  227. Dickens, D.; Rädisch, S.; Chiduza, G.N.; Giannoudis, A.; Cross, M.J.; Malik, H.; Schaeffeler, E.; Sison-Young, R.L.; Wilkinson, E.L.; Goldring, C.E.; et al. Cellular uptake of the atypical antipsychotic clozapine is a carrier-mediated process. Mol. Pharm. 2018, 15, 3557–3572. [Google Scholar] [CrossRef]
  228. Kell, D.B. The transporter-mediated cellular uptake and efflux of pharmaceutical drugs and biotechnology products: How and why phospholipid bilayer transport is negligible in real biomembranes. Molecules 2021, 26, 5629. [Google Scholar] [CrossRef] [PubMed]
  229. Hediger, M.A.; Clemencon, B.; Burrier, R.E.; Bruford, E.A. The ABCs of membrane transporters in health and disease (SLC series): Introduction. Mol. Aspects Med. 2013, 34, 95–107. [Google Scholar] [CrossRef]
  230. Anonymous. SLC Tables. Available online: http://www.bioparadigms.org/slc/intro.htm (accessed on 27 August 2019).
  231. Chen, Z.; Shi, T.; Zhang, L.; Zhu, P.; Deng, M.; Huang, C.; Hu, T.; Jiang, L.; Li, J. Mammalian drug efflux transporters of the ATP binding cassette (ABC) family in multidrug resistance: A review of the past decade. Cancer Lett. 2016, 370, 153–164. [Google Scholar] [CrossRef]
  232. Jindal, S.; Yang, L.; Day, P.J.; Kell, D.B. Involvement of multiple influx and efflux transporters in the accumulation of cationic fluorescent dyes by Escherichia coli. BMC Microbiol. 2019, 19, 195, also bioRxiv 603688v603681. [Google Scholar] [CrossRef]
  233. Ter Beek, J.; Guskov, A.; Slotboom, D.J. Structural diversity of ABC transporters. J. Gen. Physiol. 2014, 143, 419–435. [Google Scholar] [CrossRef]
  234. Lewinson, O.; Livnat-Levanon, N. Mechanism of Action of ABC Importers: Conservation, Divergence, and Physiological Adaptations. J. Mol. Biol. 2017, 429, 606–619. [Google Scholar] [CrossRef]
  235. Nigam, S.K.; Bush, K.T.; Martovetsky, G.; Ahn, S.Y.; Liu, H.C.; Richard, E.; Bhatnagar, V.; Wu, W. The organic anion transporter (OAT) family: A systems biology perspective. Physiol. Rev. 2015, 95, 83–123. [Google Scholar] [CrossRef] [PubMed]
  236. Nigam, S.K. The SLC22 Transporter Family: A Paradigm for the Impact of Drug Transporters on Metabolic Pathways, Signaling, and Disease. Annu. Rev. Pharmacol. Toxicol. 2018, 58, 663–687. [Google Scholar] [CrossRef]
  237. Bush, K.T.; Wu, W.; Lun, C.; Nigam, S.K. The drug transporter OAT3 (SLC22A8) regulates endogenous metabolite flow through the gut-liver-kidney axis. J. Biol. Chem. 2017, 292, 15789–15803. [Google Scholar] [CrossRef]
  238. Thul, P.J.; Åkesson, L.; Wiking, M.; Mahdessian, D.; Geladaki, A.; Ait Blal, H.; Alm, T.; Asplund, A.; Björk, L.; Breckels, L.M.; et al. A subcellular map of the human proteome. Science 2017, 356, eaal3321. [Google Scholar] [CrossRef]
  239. Wright Muelas, M.; Mughal, F.; O’Hagan, S.; Day, P.J.; Kell, D.B. The role and robustness of the Gini coefficient as an unbiased tool for the selection of Gini genes for normalising expression profiling data. Sci. Rep. 2019, 9, 17960. [Google Scholar] [CrossRef]
  240. Fukui, S.; Schwarcz, R.; Rapoport, S.I.; Takada, Y.; Smith, Q.R. Blood-brain barrier transport of kynurenines: Implications for brain synthesis and metabolism. J. Neurochem. 1991, 56, 2007–2017. [Google Scholar] [CrossRef] [PubMed]
  241. Sekine, A.; Kuroki, Y.; Urata, T.; Mori, N.; Fukuwatari, T. Inhibition of Large Neutral Amino Acid Transporters Suppresses Kynurenic Acid Production Via Inhibition of Kynurenine Uptake in Rodent Brain. Neurochem. Res. 2016, 41, 2256–2266. [Google Scholar] [CrossRef]
  242. Patel, W.; Rimmer, L.; Smith, M.; Moss, L.; Smith, M.A.; Snodgrass, H.R.; Pirmohamed, M.; Alfirevic, A.; Dickens, D. Probenecid Increases the Concentration of 7-Chlorokynurenic Acid Derived from the Prodrug 4-Chlorokynurenine within the Prefrontal Cortex. Mol. Pharm. 2021, 18, 113–123. [Google Scholar] [CrossRef] [PubMed]
  243. Pardridge, W.M. A Historical Review of Brain Drug Delivery. Pharmaceutics 2022, 14, 1283. [Google Scholar] [CrossRef] [PubMed]
  244. Patel, W.; Shankar, R.G.; Smith, M.A.; Snodgrass, H.R.; Pirmohamed, M.; Jorgensen, A.L.; Alfirevic, A.; Dickens, D. Role of Transporters and Enzymes in Metabolism and Distribution of 4-Chlorokynurenine (AV-101). Mol. Pharm. 2024, 21, 550–563. [Google Scholar] [CrossRef]
  245. Miller, J.M.; MacGarvey, U.; Beal, M.F. The effect of peripheral loading with kynurenine and probenecid on extracellular striatal kynurenic acid concentrations. Neurosci. Lett. 1992, 146, 115–118. [Google Scholar] [CrossRef]
  246. Vécsei, L.; Miller, J.; MacGarvey, U.; Beal, M.F. Kynurenine and probenecid inhibit pentylenetetrazol- and NMDLA-induced seizures and increase kynurenic acid concentrations in the brain. Brain Res. Bull. 1992, 28, 233–238. [Google Scholar] [CrossRef]
  247. Moroni, F.; Russi, P.; Lombardi, G.; Beni, M.; Carla, V. Presence of kynurenic acid in the mammalian brain. J. Neurochem. 1988, 51, 177–180. [Google Scholar] [CrossRef]
  248. Shepard, P.D.; Joy, B.; Clerkin, L.; Schwarcz, R. Micromolar brain levels of kynurenic acid are associated with a disruption of auditory sensory gating in the rat. Neuropsychopharmacology 2003, 28, 1454–1462. [Google Scholar] [CrossRef] [PubMed]
  249. Sas, K.; Robotka, H.; Rózsa, É.; Ágoston, M.; Szénási, G.; Gigler, G.; Marosi, M.; Kis, Z.; Farkas, T.; Vécsei, L.; et al. Kynurenine diminishes the ischemia-induced histological and electrophysiological deficits in the rat hippocampus. Neurobiol. Dis. 2008, 32, 302–308. [Google Scholar] [CrossRef]
  250. Silva-Adaya, D.; Pérez-De La Cruz, V.; Villeda-Hernández, J.; Carrillo-Mora, P.; González-Herrera, I.G.; García, E.; Colín-Barenque, L.; Pedraza-Chaverrí, J.; Santamaría, A. Protective effect of L-kynurenine and probenecid on 6-hydroxydopamine-induced striatal toxicity in rats: Implications of modulating kynurenate as a protective strategy. Neurotoxicol Teratol. 2011, 33, 303–312. [Google Scholar] [CrossRef]
  251. Russel, F.G.M.; Koenderink, J.B.; Masereeuw, R. Multidrug resistance protein 4 (MRP4/ABCC4): A versatile efflux transporter for drugs and signalling molecules. Trends Pharmacol. Sci. 2008, 29, 200–207. [Google Scholar] [CrossRef]
  252. Mendes, P.; Girardi, E.; Superti-Furga, G.; Kell, D.B. Why most transporter mutations that cause antibiotic resistance are to efflux pumps rather than to import transporters. bioRxiv 2020, 2020.2001.2016.909507v909501. [Google Scholar] [CrossRef]
  253. Li, X.-Z.; Elkins, C.A.; Zgurskaya, H.I. (Eds.) Efflux-Mediated Antimicrobial Resistance in Bacteria: Mechanisms, Regulation and Clinical Implications; Springer: Berlin/Heidelberg, Germany, 2016. [Google Scholar]
  254. Piddock, L.J.V. Clinically relevant chromosomally encoded multidrug resistance efflux pumps in bacteria. Clin. Microbiol. Rev. 2006, 19, 382–402. [Google Scholar] [CrossRef]
  255. Piddock, L.J.V. The 2019 Garrod Lecture: MDR efflux in Gram-negative bacteria-how understanding resistance led to a new tool for drug discovery. J. Antimicrob. Chemother. 2019, 74, 3128–3134. [Google Scholar] [CrossRef] [PubMed]
  256. Poku, V.O.; Iram, S.H. A critical review on modulators of Multidrug Resistance Protein 1 in cancer cells. PeerJ 2022, 10, e12594. [Google Scholar] [CrossRef]
  257. Bharathiraja, P.; Yadav, P.; Sajid, A.; Ambudkar, S.V.; Prasad, N.R. Natural medicinal compounds target signal transduction pathways to overcome ABC drug efflux transporter-mediated multidrug resistance in cancer. Drug Resist. Updat. 2023, 71, 101004. [Google Scholar] [CrossRef]
  258. Grixti, J.; O’Hagan, S.; Day, P.J.; Kell, D.B. Enhancing drug efficacy and therapeutic index through cheminformatics-based selection of small molecule binary weapons that improve transporter-mediated targeting: A cytotoxicity system based on gemcitabine. Front. Pharmacol. 2017, 8, 155. [Google Scholar] [CrossRef]
  259. Sajid, A.; Rahman, H.; Ambudkar, S.V. Advances in the structure, mechanism and targeting of chemoresistance-linked ABC transporters. Nat. Rev. Cancer 2023, 23, 762–779. [Google Scholar] [CrossRef]
  260. Wang, J.Q.; Wu, Z.X.; Yang, Y.; Teng, Q.X.; Li, Y.D.; Lei, Z.N.; Jani, K.A.; Kaushal, N.; Chen, Z.S. ATP-binding cassette (ABC) transporters in cancer: A review of recent updates. J. Evid. Based Med. 2021, 14, 232–256. [Google Scholar] [CrossRef]
  261. Boots, A.W.; Haenen, G.R.M.M.; Bast, A. Health effects of quercetin: From antioxidant to nutraceutical. Eur. J. Pharmacol. 2008, 585, 325–337. [Google Scholar] [CrossRef] [PubMed]
  262. Anand David, A.V.; Arulmoli, R.; Parasuraman, S. Overviews of Biological Importance of Quercetin: A Bioactive Flavonoid. Pharmacogn. Rev. 2016, 10, 84–89. [Google Scholar] [CrossRef] [PubMed]
  263. Mutsaers, H.A.; van den Heuvel, L.P.; Ringens, L.H.; Dankers, A.C.; Russel, F.G.; Wetzels, J.F.; Hoenderop, J.G.; Masereeuw, R. Uremic toxins inhibit transport by breast cancer resistance protein and multidrug resistance protein 4 at clinically relevant concentrations. PLoS ONE 2011, 6, e18438. [Google Scholar] [CrossRef]
  264. Dankers, A.C.A.; Mutsaers, H.A.M.; Dijkman, H.B.P.M.; van den Heuvel, L.P.; Hoenderop, J.G.; Sweep, F.C.G.J.; Russel, F.G.M.; Masereeuw, R. Hyperuricemia influences tryptophan metabolism via inhibition of multidrug resistance protein 4 (MRP4) and breast cancer resistance protein (BCRP). Biochim. Biophys. Acta 2013, 1832, 1715–1722. [Google Scholar] [CrossRef]
  265. Stafim da Cunha, R.; Azevedo, C.A.B.; Falconi, C.A.; Ruiz, F.F.; Liabeuf, S.; Carneiro-Ramos, M.S.; Stinghen, A.E.M. The Interplay between Uremic Toxins and Albumin, Membrane Transporters and Drug Interaction. Toxins 2022, 14, 177. [Google Scholar] [CrossRef]
  266. Ma, Y.; Ran, F.; Xin, M.; Gou, X.; Wang, X.; Wu, X. Albumin-bound kynurenic acid is an appropriate endogenous biomarker for assessment of the renal tubular OATs-MRP4 channel. J. Pharm. Anal. 2023, 13, 1205–1220. [Google Scholar] [CrossRef]
  267. Frechen, S.; Rostami-Hodjegan, A. Quality Assurance of PBPK Modeling Platforms and Guidance on Building, Evaluating, Verifying and Applying PBPK Models Prudently under the Umbrella of Qualification: Why, When, What, How and By Whom? Pharm. Res. 2022, 39, 1733–1748. [Google Scholar] [CrossRef]
  268. Murata, Y.; Neuhoff, S.; Rostami-Hodjegan, A.; Takita, H.; Al-Majdoub, Z.M.; Ogungbenro, K. In Vitro to In Vivo Extrapolation Linked to Physiologically Based Pharmacokinetic Models for Assessing the Brain Drug Disposition. AAPS J. 2022, 24, 28. [Google Scholar] [CrossRef]
  269. Superti-Furga, G.; Lackner, D.; Wiedmer, T.; Ingles-Prieto, A.; Barbosa, B.; Girardi, E.; Goldman, U.; Gürtl, B.; Klavins, K.; Klimek, C.; et al. The RESOLUTE consortium: Unlocking SLC transporters for drug discovery. Nat. Rev. Drug Discov. 2020, 19, 429–430. [Google Scholar] [CrossRef]
  270. Wright Muelas, M.; Roberts, I.; Mughal, F.; O’Hagan, S.; Day, P.J.; Kell, D.B. An untargeted metabolomics strategy to measure differences in metabolite uptake and excretion by mammalian cell lines. Metabolomics 2020, 16, 107. [Google Scholar] [CrossRef]
  271. Cheung, L.; Flemming, C.L.; Watt, F.; Masada, N.; Yu, D.M.; Huynh, T.; Conseil, G.; Tivnan, A.; Polinsky, A.; Gudkov, A.V.; et al. High-throughput screening identifies Ceefourin 1 and Ceefourin 2 as highly selective inhibitors of multidrug resistance protein 4 (MRP4). Biochem. Pharmacol. 2014, 91, 97–108. [Google Scholar] [CrossRef] [PubMed]
  272. Sahores, A.; Rodríguez González, A.; Yaneff, A.; May, M.; Gómez, N.; Monczor, F.; Fernández, N.; Davio, C.; Shayo, C. Ceefourin-1, a MRP4/ABCC4 inhibitor, induces apoptosis in AML cells enhanced by histamine. Biochim. Biophys. Acta Gen. Subj. 2023, 1867, 130322. [Google Scholar] [CrossRef]
  273. Marcantoni, E.; Allen, N.; Cambria, M.R.; Dann, R.; Cammer, M.; Lhakhang, T.; O’Brien, M.P.; Kim, B.; Worgall, T.; Heguy, A.; et al. Platelet Transcriptome Profiling in HIV and ATP-Binding Cassette Subfamily C Member 4 (ABCC4) as a Mediator of Platelet Activity. JACC Basic. Transl. Sci. 2018, 3, 9–22. [Google Scholar] [CrossRef] [PubMed]
  274. Rius, M.; Hummel-Eisenbeiss, J.; Keppler, D. ATP-dependent transport of leukotrienes B4 and C4 by the multidrug resistance protein ABCC4 (MRP4). J. Pharmacol. Exp. Ther. 2008, 324, 86–94. [Google Scholar] [CrossRef]
  275. Van de Ven, R.; Scheffer, G.L.; Reurs, A.W.; Lindenberg, J.J.; Oerlemans, R.; Jansen, G.; Gillet, J.P.; Glasgow, J.N.; Pereboev, A.; Curiel, D.T.; et al. A role for multidrug resistance protein 4 (MRP4; ABCC4) in human dendritic cell migration. Blood 2008, 112, 2353–2359. [Google Scholar] [CrossRef] [PubMed]
  276. Reid, G.; Wielinga, P.; Zelcer, N.; van der Heijden, I.; Kuil, A.; de Haas, M.; Wijnholds, J.; Borst, P. The human multidrug resistance protein MRP4 functions as a prostaglandin efflux transporter and is inhibited by nonsteroidal antiinflammatory drugs. Proc. Natl. Acad. Sci. USA 2003, 100, 9244–9249. [Google Scholar] [CrossRef]
  277. Gekeler, V.; Ise, W.; Sanders, K.H.; Ulrich, W.R.; Beck, J. The leukotriene LTD4 receptor antagonist MK571 specifically modulates MRP associated multidrug resistance. Biochem. Biophys. Res. Commun. 1995, 208, 345–352. [Google Scholar] [CrossRef]
  278. Takeuchi, K.; Shibata, M.; Kashiyama, E.; Umehara, K. Expression levels of multidrug resistance-associated protein 4 (MRP4) in human leukemia and lymphoma cell lines, and the inhibitory effects of the MRP-specific inhibitor MK-571 on methotrexate distribution in rats. Exp. Ther. Med. 2012, 4, 524–532. [Google Scholar] [CrossRef]
  279. Low, F.G.; Shabir, K.; Brown, J.E.; Bill, R.M.; Rothnie, A.J. Roles of ABCC1 and ABCC4 in Proliferation and Migration of Breast Cancer Cell Lines. Int. J. Mol. Sci. 2020, 21, 7664. [Google Scholar] [CrossRef]
  280. Copsel, S.; Garcia, C.; Diez, F.; Vermeulem, M.; Baldi, A.; Bianciotti, L.G.; Russel, F.G.M.; Shayo, C.; Davio, C. Multidrug resistance protein 4 (MRP4/ABCC4) regulates cAMP cellular levels and controls human leukemia cell proliferation and differentiation. J. Biol. Chem. 2011, 286, 6979–6988. [Google Scholar] [CrossRef]
  281. Britz, H.; Hanke, N.; Taub, M.E.; Wang, T.; Prasad, B.; Fernandez, E.; Stopfer, P.; Nock, V.; Lehr, T. Physiologically Based Pharmacokinetic Models of Probenecid and Furosemide to Predict Transporter Mediated Drug-Drug Interactions. Pharm. Res. 2020, 37, 250. [Google Scholar] [CrossRef]
  282. Maeda, K.; Tian, Y.; Fujita, T.; Ikeda, Y.; Kumagai, Y.; Kondo, T.; Tanabe, K.; Nakayama, H.; Horita, S.; Kusuhara, H.; et al. Inhibitory effects of p-aminohippurate and probenecid on the renal clearance of adefovir and benzylpenicillin as probe drugs for organic anion transporter (OAT) 1 and OAT3 in humans. Eur. J. Pharm. Sci. 2014, 59, 94–103. [Google Scholar] [CrossRef] [PubMed]
  283. Wu, C.P.; Calcagno, A.M.; Hladky, S.B.; Ambudkar, S.V.; Barrand, M.A. Modulatory effects of plant phenols on human multidrug-resistance proteins 1, 4 and 5 (ABCC1, 4 and 5). FEBS J. 2005, 272, 4725–4740. [Google Scholar] [CrossRef] [PubMed]
  284. Ritter, C.A.; Jedlitschky, G.; Meyer zu Schwabedissen, H.; Grube, M.; Kock, K.; Kroemer, H.K. Cellular export of drugs and signaling molecules by the ATP-binding cassette transporters MRP4 (ABCC4) and MRP5 (ABCC5). Drug Metab. Rev. 2005, 37, 253–278. [Google Scholar] [CrossRef]
  285. Wen, J.; Luo, J.; Huang, W.; Tang, J.; Zhou, H.; Zhang, W. The Pharmacological and Physiological Role of Multidrug-Resistant Protein 4. J. Pharmacol. Exp. Ther. 2015, 354, 358–375. [Google Scholar] [CrossRef]
  286. Nigam, S.K.; Granados, J.C. OAT, OATP, and MRP Drug Transporters and the Remote Sensing and Signaling Theory. Annu. Rev. Pharmacol. Toxicol. 2023, 63, 637–660. [Google Scholar] [CrossRef]
  287. El-Sheikh, A.A.K.; van den Heuvel, J.J.M.W.; Koenderink, J.B.; Russel, F.G.M. Interaction of nonsteroidal anti-inflammatory drugs with multidrug resistance protein (MRP) 2/ABCC2- and MRP4/ABCC4-mediated methotrexate transport. J. Pharmacol. Exp. Ther. 2007, 320, 229–235. [Google Scholar] [CrossRef] [PubMed]
  288. Bartlett, R.D.; Esslinger, C.S.; Thompson, C.M.; Bridges, R.J. Substituted quinolines as inhibitors of L-glutamate transport into synaptic vesicles. Neuropharmacology 1998, 37, 839–846. [Google Scholar] [CrossRef]
  289. Kanai, Y.; Clemencon, B.; Simonin, A.; Leuenberger, M.; Lochner, M.; Weisstanner, M.; Hediger, M.A. The SLC1 high-affinity glutamate and neutral amino acid transporter family. Mol. Aspects Med. 2013, 34, 108–120. [Google Scholar] [CrossRef]
  290. Magi, S.; Piccirillo, S.; Amoroso, S.; Lariccia, V. Excitatory Amino Acid Transporters (EAATs): Glutamate Transport and Beyond. Int. J. Mol. Sci. 2019, 20, 5674. [Google Scholar] [CrossRef] [PubMed]
  291. Halestrap, A.P. The SLC16 gene family—Structure, role and regulation in health and disease. Mol. Aspects Med. 2013, 34, 337–349. [Google Scholar] [CrossRef]
  292. Felmlee, M.A.; Jones, R.S.; Rodriguez-Cruz, V.; Follman, K.E.; Morris, M.E. Monocarboxylate Transporters (SLC16): Function, Regulation, and Role in Health and Disease. Pharmacol. Rev. 2020, 72, 466–485. [Google Scholar] [CrossRef]
  293. Bosshart, P.D.; Charles, R.P.; Garibsingh, R.A.; Schlessinger, A.; Fotiadis, D. SLC16 Family: From Atomic Structure to Human Disease. Trends Biochem. Sci. 2021, 46, 28–40. [Google Scholar] [CrossRef]
  294. Tan, K.M.; Tint, M.T.; Kothandaraman, N.; Michael, N.; Sadananthan, S.A.; Velan, S.S.; Fortier, M.V.; Yap, F.; Tan, K.H.; Gluckman, P.D.; et al. The Kynurenine Pathway Metabolites in Cord Blood Positively Correlate With Early Childhood Adiposity. J. Clin. Endocrinol. Metab. 2022, 107, e2464–e2473. [Google Scholar] [CrossRef] [PubMed]
  295. Notarangelo, F.M.; Schwarcz, R. Restraint Stress during Pregnancy Rapidly Raises Kynurenic Acid Levels in Mouse Placenta and Fetal Brain. Dev. Neurosci. 2016, 38, 458–468. [Google Scholar] [CrossRef]
  296. Goeden, N.; Notarangelo, F.M.; Pocivavsek, A.; Beggiato, S.; Bonnin, A.; Schwarcz, R. Prenatal Dynamics of Kynurenine Pathway Metabolism in Mice: Focus on Kynurenic Acid. Dev. Neurosci. 2017, 39, 519–528. [Google Scholar] [CrossRef]
  297. Lin, L.; Lemieux, G.A.; Enogieru, O.J.; Giacomini, K.M.; Ashrafi, K. Neural production of kynurenic acid in Caenorhabditis elegans requires the AAT-1 transporter. Genes Dev. 2020, 34, 1033–1038. [Google Scholar] [CrossRef] [PubMed]
  298. Scalise, M.; Galluccio, M.; Console, L.; Pochini, L.; Indiveri, C. The Human SLC7A5 (LAT1): The Intriguing Histidine/Large Neutral Amino Acid Transporter and Its Relevance to Human Health. Front. Chem. 2018, 6, 243. [Google Scholar] [CrossRef]
  299. Cappoli, N.; Jenkinson, M.D.; Dello Russo, C.; Dickens, D. LAT1, a novel pharmacological target for the treatment of glioblastoma. Biochem. Pharmacol. 2022, 201, 115103. [Google Scholar] [CrossRef]
  300. Kanai, Y. Amino acid transporter LAT1 (SLC7A5) as a molecular target for cancer diagnosis and therapeutics. Pharmacol. Ther. 2022, 230, 107964. [Google Scholar] [CrossRef] [PubMed]
  301. Nishikubo, K.; Ohgaki, R.; Okanishi, H.; Okuda, S.; Xu, M.; Endou, H.; Kanai, Y. Pharmacologic inhibition of LAT1 predominantly suppresses transport of large neutral amino acids and downregulates global translation in cancer cells. J. Cell Mol. Med. 2022, 26, 5246–5256. [Google Scholar] [CrossRef] [PubMed]
  302. Sekine, A.; Okamoto, M.; Kanatani, Y.; Sano, M.; Shibata, K.; Fukuwatari, T. Amino acids inhibit kynurenic acid formation via suppression of kynurenine uptake or kynurenic acid synthesis in rat brain in vitro. Springerplus 2015, 4, 48. [Google Scholar] [CrossRef] [PubMed]
  303. Fukushima, T.; Sone, Y.; Mitsuhashi, S.; Tomiya, M.; Toyo’oka, T. Alteration of kynurenic acid concentration in rat plasma following optically pure kynurenine administration: A comparative study between enantiomers. Chirality 2009, 21, 468–472. [Google Scholar] [CrossRef] [PubMed]
  304. Pérez-de la Cruz, V.; Amori, L.; Sathyasaikumar, K.V.; Wang, X.D.; Notarangelo, F.M.; Wu, H.Q.; Schwarcz, R. Enzymatic transamination of D-kynurenine generates kynurenic acid in rat and human brain. J. Neurochem. 2012, 120, 1026–1035. [Google Scholar] [CrossRef] [PubMed]
  305. Takahashi, H.; Kaihara, M.; Price, J.M. The conversion of kynurenic acid to quinaldic acid by humans and rats. J. Biol. Chem. 1956, 223, 705–708. [Google Scholar]
  306. Turska, M.; Rutyna, R.; Paluszkiewicz, M.; Terlecka, P.; Dobrowolski, A.; Pelak, J.; Turski, M.P.; Muszynska, B.; Dabrowski, W.; Kocki, T.; et al. Presence of kynurenic acid in alcoholic beverages—Is this good news, or bad news? Med. Hypotheses 2019, 122, 200–205. [Google Scholar] [CrossRef] [PubMed]
  307. Sadok, I.; Jędruchniewicz, K. Dietary Kynurenine Pathway Metabolites-Source, Fate, and Chromatographic Determinations. Int. J. Mol. Sci. 2023, 24, 16304. [Google Scholar] [CrossRef] [PubMed]
  308. Turska, M.; Paluszkiewicz, P.; Turski, W.A.; Parada-Turska, J. A Review of the Health Benefits of Food Enriched with Kynurenic Acid. Nutrients 2022, 14, 4182. [Google Scholar] [CrossRef] [PubMed]
  309. Varga, N.; Csapó, E.; Majláth, Z.; Ilisz, I.; Krizbai, I.A.; Wilhelm, I.; Knapp, L.; Toldi, J.; Vécsei, L.; Dékány, I. Targeting of the kynurenic acid across the blood-brain barrier by core-shell nanoparticles. Eur. J. Pharm. Sci. 2016, 86, 67–74. [Google Scholar] [CrossRef] [PubMed]
  310. O’Hagan, S.; Kell, D.B. The apparent permeabilities of Caco-2 cells to marketed drugs: Magnitude, and independence from both biophysical properties and endogenite similarities. PeerJ 2015, 3, e1405. [Google Scholar] [PubMed]
  311. Füvesi, J.; Somlai, C.; Németh, H.; Varga, H.; Kis, Z.; Farkas, T.; Károly, N.; Dobszay, M.; Penke, Z.; Penke, B.; et al. Comparative study on the effects of kynurenic acid and glucosamine-kynurenic acid. Pharmacol. Biochem. Behav. 2004, 77, 95–102. [Google Scholar] [CrossRef] [PubMed]
  312. Robotka, H.; Németh, H.; Somlai, C.; Vécsei, L.; Toldi, J. Systemically administered glucosamine-kynurenic acid, but not pure kynurenic acid, is effective in decreasing the evoked activity in area CA1 of the rat hippocampus. Eur. J. Pharmacol. 2005, 513, 75–80. [Google Scholar] [CrossRef] [PubMed]
  313. Hornok, V.; Bujdosó, T.; Toldi, J.; Nagy, K.; Demeter, I.; Fazakas, C.; Krizbai, I.; Vécsei, L.; Dékány, I. Preparation and properties of nanoscale containers for biomedical application in drug delivery: Preliminary studies with kynurenic acid. J. Neural Transm. 2012, 119, 115–121. [Google Scholar] [CrossRef] [PubMed]
  314. Hornok, V.; Amin, K.W.K.; Kovács, A.N.; Juhász, Á.; Katona, G.; Balogh, G.T.; Csapó, E. Increased blood-brain barrier permeability of neuroprotective drug by colloidal serum albumin carriers. Colloids Surf. B Biointerfaces 2022, 220, 112935. [Google Scholar] [CrossRef]
  315. Juhász, A.; Ungor, D.; Varga, N.; Katona, G.; Balogh, G.T.; Csapó, E. Lipid-Based Nanocarriers for Delivery of Neuroprotective Kynurenic Acid: Preparation, Characterization, and BBB Transport. Int. J. Mol. Sci. 2023, 24, 14251. [Google Scholar] [CrossRef] [PubMed]
  316. Monfared, Y.K.; Pedrazzo, A.R.; Mahmoudian, M.; Caldera, F.; Zakeri-Milani, P.; Valizadeh, H.; Cavalli, R.; Matencio, A.; Trotta, F. Oral supplementation of solvent-free kynurenic acid/cyclodextrin nanosponges complexes increased its bioavailability. Colloids Surf. B Biointerfaces 2023, 222, 113101. [Google Scholar] [CrossRef]
  317. Deák, Á.; Csapó, E.; Juhász, Á.; Dékány, I.; Janovák, L. Anti-ulcerant kynurenic acid molecules intercalated Mg/Al-layered double hydroxide and its release study. Appl. Clay Sci. 2018, 156, 28–35. [Google Scholar] [CrossRef]
  318. Dhakar, N.K.; Caldera, F.; Bessone, F.; Cecone, C.; Pedrazzo, A.R.; Cavalli, R.; Dianzani, C.; Trotta, F. Evaluation of solubility enhancement, antioxidant activity, and cytotoxicity studies of kynurenic acid loaded cyclodextrin nanosponge. Carbohydr. Polym. 2019, 224, 115168. [Google Scholar] [CrossRef] [PubMed]
  319. Cheah, I.K.; Tang, R.M.Y.; Yew, T.S.Z.; Lim, K.H.C.; Halliwell, B. Administration of Pure Ergothioneine to Healthy Human Subjects: Uptake, Metabolism, and Effects on Biomarkers of Oxidative Damage and Inflammation. Antioxid. Redox Signal. 2017, 26, 193–206. [Google Scholar] [CrossRef] [PubMed]
  320. Kenny, L.C.; The SCOPE Consortium; Brown, L.W.; Ortea, P.; Tuytten, R.; Kell, D.B. Relationship between the concentration of ergothioneine in plasma and the likelihood of developing pre-eclampsia. Biosci. Rep. 2023, 43, BSR20230160. [Google Scholar] [CrossRef]
  321. Uwai, Y.; Honjo, H.; Iwamoto, K. Interaction and transport of kynurenic acid via human organic anion transporters hOAT1 and hOAT3. Pharmacol. Res. 2012, 65, 254–260. [Google Scholar] [CrossRef] [PubMed]
  322. Uwai, Y.; Hara, H.; Iwamoto, K. Transport of Kynurenic Acid by Rat Organic Anion Transporters rOAT1 and rOAT3: Species Difference between Human and Rat in OAT1. Int. J. Tryptophan Res. 2013, 6, 1–6. [Google Scholar] [CrossRef] [PubMed]
  323. Stone, T.W. Neuropharmacology of quinolinic and kynurenic acids. Pharmacol. Rev. 1993, 45, 309–379. [Google Scholar] [PubMed]
  324. Kell, D.B.; Pretorius, E. Serum ferritin is an important disease marker, and is mainly a leakage product from damaged cells. Metallomics 2014, 6, 748–773. [Google Scholar] [CrossRef] [PubMed]
  325. Liu, J.; Xi, K.; Zhang, L.; Han, M.; Wang, Q.; Liu, X. Tryptophan metabolites and gut microbiota play an important role in pediatric migraine diagnosis. J. Headache Pain 2024, 25, 2. [Google Scholar] [CrossRef] [PubMed]
  326. Keszthelyi, D.; Troost, F.J.; Jonkers, D.M.; Kruimel, J.W.; Leue, C.; Masclee, A.A. Decreased levels of kynurenic acid in the intestinal mucosa of IBS patients: Relation to serotonin and psychological state. J. Psychosom. Res. 2013, 74, 501–504. [Google Scholar] [CrossRef] [PubMed]
  327. Bakker, L.; Choe, K.; Eussen, S.; Ramakers, I.; van den Hove, D.L.A.; Kenis, G.; Rutten, B.P.F.; Verhey, F.R.J.; Kohler, S. Relation of the kynurenine pathway with normal age: A systematic review. Mech. Ageing Dev. 2024, 217, 111890. [Google Scholar] [CrossRef] [PubMed]
  328. Sorgdrager, F.J.H.; Vermeiren, Y.; Van Faassen, M.; van der Ley, C.; Nollen, E.A.A.; Kema, I.P.; De Deyn, P.P. Age- and disease-specific changes of the kynurenine pathway in Parkinson’s and Alzheimer’s disease. J. Neurochem. 2019, 151, 656–668. [Google Scholar] [CrossRef] [PubMed]
  329. Westbrook, R.; Chung, T.; Lovett, J.; Ward, C.; Joca, H.; Yang, H.; Khadeer, M.; Tian, J.; Xue, Q.L.; Le, A.; et al. Kynurenines link chronic inflammation to functional decline and physical frailty. JCI Insight 2020, 5, e136091. [Google Scholar] [CrossRef]
  330. Braidy, N.; Guillemin, G.J.; Mansour, H.; Chan-Ling, T.; Grant, R. Changes in kynurenine pathway metabolism in the brain, liver and kidney of aged female Wistar rats. FEBS J. 2011, 278, 4425–4434. [Google Scholar] [CrossRef] [PubMed]
  331. Roberts, I.; Wright Muelas, M.; Taylor, J.M.; Davison, A.S.; Winder, C.L.; Goodacre, R.; Kell, D.B. Quantitative LC-MS study of compounds found predictive of COVID-19 severity and outcome. Metabolomics 2023, 19, 87. [Google Scholar] [CrossRef]
  332. Theiler-Schwetz, V.; Trummer, C.; Grubler, M.R.; Keppel, M.H.; Zittermann, A.; Tomaschitz, A.; Marz, W.; Meinitzer, A.; Pilz, S. Associations of Parameters of the Tryptophan-Kynurenine Pathway with Cardiovascular Risk Factors in Hypertensive Patients. Nutrients 2023, 15, 256. [Google Scholar] [CrossRef]
  333. Aarsland, T.I.; Landaas, E.T.; Hegvik, T.A.; Ulvik, A.; Halmoy, A.; Ueland, P.M.; Haavik, J. Serum concentrations of kynurenines in adult patients with attention-deficit hyperactivity disorder (ADHD): A case-control study. Behav. Brain Funct. 2015, 11, 36. [Google Scholar] [CrossRef]
  334. Dudzińska, E.; Szymona, K.; Kloc, R.; Gil-Kulik, P.; Kocki, T.; Świstowska, M.; Bogucki, J.; Kocki, J.; Urbanska, E.M. Increased expression of kynurenine aminotransferases mRNA in lymphocytes of patients with inflammatory bowel disease. Therap Adv. Gastroenterol. 2019, 12, 1756284819881304. [Google Scholar] [CrossRef]
  335. Fukushima, T.; Mitsuhashi, S.; Tomiya, M.; Iyo, M.; Hashimoto, K.; Toyo’oka, T. Determination of kynurenic acid in human serum and its correlation with the concentration of certain amino acids. Clin. Chim. Acta 2007, 377, 174–178. [Google Scholar] [CrossRef] [PubMed]
  336. Iłźecka, J.; Kocki, T.; Stelmasiak, Z.; Turski, W.A. Endogenous protectant kynurenic acid in amyotrophic lateral sclerosis. Acta Neurol. Scand. 2003, 107, 412–418. [Google Scholar] [CrossRef]
  337. Zhao, Y.J.; Zhou, C.; Wei, Y.Y.; Li, H.H.; Lei, W.; Boeldt, D.S.; Wang, K.; Zheng, J. Differential Distribution of Tryptophan-Metabolites in Fetal and Maternal Circulations During Normotensive and Preeclamptic Pregnancies. Reprod. Sci. 2022, 29, 1278–1286. [Google Scholar] [CrossRef] [PubMed]
  338. Hartai, Z.; Klivenyi, P.; Janaky, T.; Penke, B.; Dux, L.; Vecsei, L. Kynurenine metabolism in plasma and in red blood cells in Parkinson’s disease. J. Neurol. Sci. 2005, 239, 31–35. [Google Scholar] [CrossRef]
  339. Hu, L.J.; Li, X.F.; Hu, J.Q.; Ni, X.J.; Lu, H.Y.; Wang, J.J.; Huang, X.N.; Lin, C.X.; Shang, D.W.; Wen, Y.G. A Simple HPLC-MS/MS Method for Determination of Tryptophan, Kynurenine and Kynurenic Acid in Human Serum and its Potential for Monitoring Antidepressant Therapy. J. Anal. Toxicol. 2017, 41, 37–44. [Google Scholar] [CrossRef]
  340. Özdemir, M.; Abusoglu, S.; Baldane, S.; Kıraç, C.O.; Unlu, A.; Onmaz, D.E.; Çelik, M.; Abusoglu, G. Analyzing serum tryptophan metabolites in patients with gestational diabetes mellitus. Rev. Rom. Med. Lab. 2023, 31, 251–262. [Google Scholar] [CrossRef]
  341. Murthi, P.; Wallace, E.M.; Walker, D.W. Altered placental tryptophan metabolic pathway in human fetal growth restriction. Placenta 2017, 52, 62–70. [Google Scholar] [CrossRef] [PubMed]
  342. Tuka, B.; Nyári, A.; Cseh, E.K.; Körtési, T.; Veréb, D.; Tömösi, F.; Kecskeméti, G.; Janáky, T.; Tajti, J.; Vécsei, L. Clinical relevance of depressed kynurenine pathway in episodic migraine patients: Potential prognostic markers in the peripheral plasma during the interictal period. J. Headache Pain 2021, 22, 60. [Google Scholar] [CrossRef] [PubMed]
  343. Gulaj, E.; Pawlak, K.; Bien, B.; Pawlak, D. Kynurenine and its metabolites in Alzheimer’s disease patients. Adv. Med. Sci. 2010, 55, 204–211. [Google Scholar] [CrossRef]
  344. Savitz, J. Blood versus cerebrospinal fluid: Kynurenine pathway metabolites in depression. Brain Behav. Immun. 2022, 101, 333–334. [Google Scholar] [CrossRef]
  345. Savitz, J.; Ford, B.N.; Kuplicki, R.; Khalsa, S.; Teague, T.K.; Paulus, M.P. Acute administration of ibuprofen increases serum concentration of the neuroprotective kynurenine pathway metabolite, kynurenic acid: A pilot randomized, placebo-controlled, crossover study. Psychopharmacology 2022, 239, 3919–3927. [Google Scholar] [CrossRef]
  346. Tuka, B.; Körtési, T.; Nánási, N.; Tömösi, F.; Janáky, T.; Veréb, D.; Szok, D.; Tajti, J.; Vécsei, L. Cluster headache and kynurenines. J. Headache Pain 2023, 24, 35. [Google Scholar] [CrossRef]
  347. Sellgren, C.M.; Gracias, J.; Jungholm, O.; Perlis, R.H.; Engberg, G.; Schwieler, L.; Landen, M.; Erhardt, S. Peripheral and central levels of kynurenic acid in bipolar disorder subjects and healthy controls. Transl. Psychiatry 2019, 9, 37. [Google Scholar] [CrossRef]
  348. Apalset, E.M.; Gjesdal, C.G.; Ueland, P.M.; Midttun, O.; Ulvik, A.; Eide, G.E.; Meyer, K.; Tell, G.S. Interferon (IFN)-gamma-mediated inflammation and the kynurenine pathway in relation to bone mineral density: The Hordaland Health Study. Clin. Exp. Immunol. 2014, 176, 452–460. [Google Scholar] [CrossRef]
  349. Tang, J.; Shen, H.; Zhao, X.; Holenarsipur, V.K.; Mariappan, T.T.; Zhang, Y.; Panfen, E.; Zheng, J.; Humphreys, W.G.; Lai, Y. Endogenous Plasma Kynurenic Acid in Human: A Newly Discovered Biomarker for Drug-Drug Interactions Involving Organic Anion Transporter 1 and 3 Inhibition. Drug Metab. Dispos. 2021, 49, 1063–1069. [Google Scholar] [CrossRef] [PubMed]
  350. Turski, M.P.; Turska, M.; Paluszkiewicz, P.; Parada-Turska, J.; Oxenkrug, G.F. Kynurenic Acid in the digestive system-new facts, new challenges. Int. J. Tryptophan Res. 2013, 6, 47–55. [Google Scholar] [CrossRef] [PubMed]
  351. Amirkhani, A.; Heldin, E.; Markides, K.E.; Bergquist, J. Quantitation of tryptophan, kynurenine and kynurenic acid in human plasma by capillary liquid chromatography-electrospray ionization tandem mass spectrometry. J. Chromatogr. B Analyt Technol. Biomed. Life Sci. 2002, 780, 381–387. [Google Scholar] [CrossRef] [PubMed]
  352. Zhao, J.; Gao, P.; Zhu, D. Optimization of Zn2+-containing mobile phase for simultaneous determination of kynurenine, kynurenic acid and tryptophan in human plasma by high performance liquid chromatography. J. Chromatogr. B Analyt Technol. Biomed. Life Sci. 2010, 878, 603–608. [Google Scholar] [CrossRef] [PubMed]
  353. Butler, M.I.; Long-Smith, C.; Moloney, G.M.; Morkl, S.; O’Mahony, S.M.; Cryan, J.F.; Clarke, G.; Dinan, T.G. The immune-kynurenine pathway in social anxiety disorder. Brain Behav. Immun. 2022, 99, 317–326. [Google Scholar] [CrossRef] [PubMed]
  354. Nilsen, R.M.; Bjørke-Monsen, A.L.; Midttun, Ø.; Nygård, O.; Pedersen, E.R.; Ulvik, A.; Magnus, P.; Gjessing, H.K.; Vollset, S.E.; Ueland, P.M. Maternal tryptophan and kynurenine pathway metabolites and risk of preeclampsia. Obstet. Gynecol. 2012, 119, 1243–1250. [Google Scholar] [CrossRef]
  355. González-Sánchez, M.; Jiménez, J.; Narváez, A.; Antequera, D.; Llamas-Velasco, S.; Martín, A.H.; Arjona, J.A.M.; López de Munain, A.; Bisa, A.L.; Marco, M.P.; et al. Kynurenic Acid Levels are Increased in the CSF of Alzheimer’s Disease Patients. Biomolecules 2020, 10, 571. [Google Scholar] [CrossRef] [PubMed]
  356. Swartz, K.J.; Matson, W.R.; MacGarvey, U.; Ryan, E.A.; Beal, M.F. Measurement of kynurenic acid in mammalian brain extracts and cerebrospinal fluid by high-performance liquid chromatography with fluorometric and coulometric electrode array detection. Anal. Biochem. 1990, 185, 363–376. [Google Scholar] [CrossRef] [PubMed]
  357. Kepplinger, B.; Baran, H.; Kainz, A.; Ferraz-Leite, H.; Newcombe, J.; Kalina, P. Age-related increase of kynurenic acid in human cerebrospinal fluid-IgG and beta2-microglobulin changes. Neurosignals 2005, 14, 126–135. [Google Scholar] [CrossRef] [PubMed]
  358. Coutinho, L.G.; Christen, S.; Bellac, C.L.; Fontes, F.L.; Souza, F.R.; Grandgirard, D.; Leib, S.L.; Agnez-Lima, L.F. The kynurenine pathway is involved in bacterial meningitis. J. Neuroinflammation 2014, 11, 169. [Google Scholar] [CrossRef] [PubMed]
  359. Milart, P.; Paluszkiewicz, P.; Dobrowolski, P.; Tomaszewska, E.; Smolinska, K.; Debinska, I.; Gawel, K.; Walczak, K.; Bednarski, J.; Turska, M.; et al. Kynurenic acid as the neglected ingredient of commercial baby formulas. Sci. Rep. 2019, 9, 6108. [Google Scholar] [CrossRef]
  360. Li, J.; Zhang, Y.; Yang, S.; Lu, Z.; Li, G.; Wu, S.; Wu, D.R.; Liu, J.; Zhou, B.; Wang, H.D.; et al. The Beneficial Effects of Edible Kynurenic Acid from Marine Horseshoe Crab (Tachypleus tridentatus) on Obesity, Hyperlipidemia, and Gut Microbiota in High-Fat Diet-Fed Mice. Oxid. Med. Cell Longev. 2021, 2021, 8874503. [Google Scholar] [CrossRef]
  361. Tomaszewska, E.; Muszynski, S.; Kuc, D.; Dobrowolski, P.; Lamorski, K.; Smolinska, K.; Donaldson, J.; Swietlicka, I.; Mielnik-Blaszczak, M.; Paluszkiewicz, P.; et al. Chronic dietary supplementation with kynurenic acid, a neuroactive metabolite of tryptophan, decreased body weight without negative influence on densitometry and mandibular bone biomechanical endurance in young rats. PLoS ONE 2019, 14, e0226205. [Google Scholar] [CrossRef]
  362. Paluszkiewicz, P.; Zgrajka, W.; Saran, T.; Schabowski, J.; Piedra, J.L.; Fedkiv, O.; Rengman, S.; Pierzynowski, S.G.; Turski, W.A. High concentration of kynurenic acid in bile and pancreatic juice. Amino Acids 2009, 37, 637–641. [Google Scholar] [CrossRef]
  363. Dehhaghi, M.; Kazemi Shariat Panahi, H.; Heng, B.; Guillemin, G.J. The Gut Microbiota, Kynurenine Pathway, and Immune System Interaction in the Development of Brain Cancer. Front. Cell Dev. Biol. 2020, 8, 562812. [Google Scholar] [CrossRef]
  364. Arsenescu, R.; Arsenescu, V.; Zhong, J.; Nasser, M.; Melinte, R.; Dingle, R.W.; Swanson, H.; de Villiers, W.J. Role of the xenobiotic receptor in inflammatory bowel disease. Inflamm. Bowel Dis. 2011, 17, 1149–1162. [Google Scholar] [CrossRef] [PubMed]
  365. Benson, J.M.; Shepherd, D.M. Aryl hydrocarbon receptor activation by TCDD reduces inflammation associated with Crohn’s disease. Toxicol. Sci. 2011, 120, 68–78. [Google Scholar] [CrossRef] [PubMed]
  366. Korecka, A.; Dona, A.; Lahiri, S.; Tett, A.J.; Al-Asmakh, M.; Braniste, V.; D’Arienzo, R.; Abbaspour, A.; Reichardt, N.; Fujii-Kuriyama, Y.; et al. Bidirectional communication between the Aryl hydrocarbon Receptor (AhR) and the microbiome tunes host metabolism. NPJ Biofilms Microbiomes 2016, 2, 16014. [Google Scholar] [CrossRef]
  367. Qiu, J.; Heller, J.J.; Guo, X.; Chen, Z.M.; Fish, K.; Fu, Y.X.; Zhou, L. The aryl hydrocarbon receptor regulates gut immunity through modulation of innate lymphoid cells. Immunity 2012, 36, 92–104. [Google Scholar] [CrossRef]
  368. Shi, L.Z.; Faith, N.G.; Nakayama, Y.; Suresh, M.; Steinberg, H.; Czuprynski, C.J. The aryl hydrocarbon receptor is required for optimal resistance to Listeria monocytogenes infection in mice. J. Immunol. 2007, 179, 6952–6962. [Google Scholar] [CrossRef]
  369. Sutter, C.H.; Bodreddigari, S.; Campion, C.; Wible, R.S.; Sutter, T.R. 2,3,7,8-Tetrachlorodibenzo-p-dioxin increases the expression of genes in the human epidermal differentiation complex and accelerates epidermal barrier formation. Toxicol. Sci. 2011, 124, 128–137. [Google Scholar] [CrossRef]
  370. Cheong, J.E.; Sun, L. Targeting the IDO1/TDO2-KYN-AhR Pathway for Cancer Immunotherapy-Challenges and Opportunities. Trends Pharmacol. Sci. 2018, 39, 307–325. [Google Scholar] [CrossRef]
  371. Dong, F.; Hao, F.; Murray, I.A.; Smith, P.B.; Koo, I.; Tindall, A.M.; Kris-Etherton, P.M.; Gowda, K.; Amin, S.G.; Patterson, A.D.; et al. Intestinal microbiota-derived tryptophan metabolites are predictive of Ah receptor activity. Gut Microbes 2020, 12, 1–24. [Google Scholar] [CrossRef]
  372. Anzai, H.; Yoshimoto, S.; Okamura, K.; Hiraki, A.; Hashimoto, S. IDO1-mediated Trp-kynurenine-AhR signal activation induces stemness and tumor dormancy in oral squamous cell carcinomas. Oral. Sci. Internat 2022, 19, 31–43. [Google Scholar] [CrossRef]
  373. Gao, J.; Xu, K.; Liu, H.; Liu, G.; Bai, M.; Peng, C.; Li, T.; Yin, Y. Impact of the Gut Microbiota on Intestinal Immunity Mediated by Tryptophan Metabolism. Front. Cell Infect. Microbiol. 2018, 8, 13. [Google Scholar] [CrossRef]
  374. Aleti, G.; Troyer, E.A.; Hong, S. G protein-coupled receptors: A target for microbial metabolites and a mechanistic link to microbiome-immune-brain interactions. Brain Behav. Immun. Health 2023, 32, 100671. [Google Scholar] [CrossRef]
  375. Sanders, K.M. G protein-coupled receptors in gastrointestinal physiology. IV. Neural regulation of gastrointestinal smooth muscle. Am. J. Physiol. 1998, 275, G1-7. [Google Scholar] [CrossRef] [PubMed]
  376. Zeng, Z.; Ma, C.; Chen, K.; Jiang, M.; Vasu, R.; Liu, R.; Zhao, Y.; Zhang, H. Roles of G Protein-Coupled Receptors (GPCRs) in Gastrointestinal Cancers: Focus on Sphingosine 1-Shosphate Receptors, Angiotensin II Receptors, and Estrogen-Related GPCRs. Cells 2021, 10, 2988. [Google Scholar] [CrossRef]
  377. Wang, D.; Wang, W.; Bing, X.; Xu, C.; Qiu, J.; Shen, J.; Huang, J.; Li, J.; Liu, P.; Xie, B. GPR35-mediated kynurenic acid sensing contributes to maintenance of gut microbiota homeostasis in ulcerative colitis. FEBS Open Bio 2023, 13, 1415–1433. [Google Scholar] [CrossRef] [PubMed]
  378. Wang, D.; Li, D.; Zhang, Y.; Chen, J.; Zhang, Y.; Liao, C.; Qin, S.; Tian, Y.; Zhang, Z.; Xu, F. Functional metabolomics reveal the role of AHR/GPR35 mediated kynurenic acid gradient sensing in chemotherapy-induced intestinal damage. Acta Pharm. Sin. B 2021, 11, 763–780. [Google Scholar] [CrossRef] [PubMed]
  379. Miyamoto, K.; Sujino, T.; Harada, Y.; Ashida, H.; Yoshimatsu, Y.; Yonemoto, Y.; Nemoto, Y.; Tomura, M.; Melhem, H.; Niess, J.H.; et al. The gut microbiota-induced kynurenic acid recruits GPR35-positive macrophages to promote experimental encephalitis. Cell Rep. 2023, 42, 113005. [Google Scholar] [CrossRef]
  380. Pedraz-Petrozzi, B.; Marszalek-Grabska, M.; Kozub, A.; Szalaj, K.; Trzpil, A.; Stachniuk, A.; Lamade, E.K.; Gilles, M.; Deuschle, M.; Turski, W.A.; et al. LC-MS/MS-based quantification of tryptophan, kynurenine, and kynurenic acid in human placental, fetal membranes, and umbilical cord samples. Sci. Rep. 2023, 13, 12554. [Google Scholar] [CrossRef]
  381. Yan, J.; Kuzhiumparambil, U.; Bandodkar, S.; Solowij, N.; Fu, S. Development and validation of a simple, rapid and sensitive LC-MS/MS method for the measurement of urinary neurotransmitters and their metabolites. Anal. Bioanal. Chem. 2017, 409, 7191–7199. [Google Scholar] [CrossRef] [PubMed]
  382. Oluwagbemigun, K.; Anesi, A.; Clarke, G.; Schmid, M.; Mattivi, F.; Nothlings, U. An Investigation into the Temporal Reproducibility of Tryptophan Metabolite Networks Among Healthy Adolescents. Int. J. Tryptophan Res. 2021, 14, 11786469211041376. [Google Scholar] [CrossRef]
  383. Pritchard, J.B.; Miller, D.S. Mechanisms mediating renal secretion of organic anions and cations. Physiol. Rev. 1993, 73, 765–796. [Google Scholar] [CrossRef]
  384. Sekine, T.; Cha, S.H.; Endou, H. The multispecific organic anion transporter (OAT) family. Pflugers Arch. 2000, 440, 337–350. [Google Scholar] [CrossRef] [PubMed]
  385. Shi, Y.; Luo, S.; Zhai, J.; Chen, Y. A novel causative role of imbalanced kynurenine pathway in ulcerative colitis: Upregulation of KMO and KYNU promotes intestinal inflammation. Biochim. Biophys. Acta Mol. Basis Dis. 2024, 1870, 166929. [Google Scholar] [CrossRef]
  386. Turski, M.P.; Turska, M.; Zgrajka, W.; Bartnik, M.; Kocki, T.; Turski, W.A. Distribution, synthesis, and absorption of kynurenic acid in plants. Planta Med. 2011, 77, 858–864. [Google Scholar] [CrossRef]
  387. Starratt, A.N.; Caveney, S. Quinoline-2-carboxylic acids from Ephedra species. Phytochemistry 1996, 42, 1477–1478. [Google Scholar] [CrossRef]
  388. Michael, J.P. Quinoline, quinazoline and acridone alkaloids. Nat. Prod. Rep. 1998, 15, 595–606. [Google Scholar]
  389. Zgrajka, W.; Turska, M.; Rajtar, G.; Majdan, M.; Parada-Turska, J. Kynurenic acid content in anti-rheumatic herbs. Ann. Agric. Environ. Med. 2013, 20, 800–802. [Google Scholar] [PubMed]
  390. Catchpole, G.S.; Beckmann, M.; Enot, D.P.; Mondhe, M.; Zywicki, B.; Taylor, J.; Hardy, N.; Smith, A.; King, R.D.; Kell, D.B.; et al. Hierarchical metabolomics demonstrates substantial compositional similarity between genetically modified and conventional potato crops. Proc. Natl. Acad. Sci. USA 2005, 102, 14458–14462. [Google Scholar] [PubMed]
  391. Turski, M.P.; Kaminski, P.; Zgrajka, W.; Turska, M.; Turski, W.A. Potato- an important source of nutritional kynurenic acid. Plant Foods Hum. Nutr. 2012, 67, 17–23. [Google Scholar] [CrossRef] [PubMed]
  392. Turski, M.P.; Chwil, S.; Turska, M.; Chwil, M.; Kocki, T.; Rajtar, G.; Parada-Turska, J. An exceptionally high content of kynurenic acid in chestnut honey and flowers of chestnut tree. J. Food Composition Anal. 2016, 48, 67–72. [Google Scholar] [CrossRef]
  393. Truchado, P.; Martos, I.; Bortolotti, L.; Sabatini, A.G.; Ferreres, F.; Tomas-Barberan, F.A. Use of quinoline alkaloids as markers of the floral origin of chestnut honey. J. Agric. Food Chem. 2009, 57, 5680–5686. [Google Scholar] [CrossRef] [PubMed]
  394. Greenaway, W.; May, J.; Scaysbrook, T.; Whatley, F.R. Identification by gas chromatography-mass spectrometry of 150 compounds in propolis. Z. Naturforsch. C 1991, 46, 111–121. [Google Scholar]
  395. Toreti, V.C.; Sato, H.H.; Pastore, G.M.; Park, Y.K. Recent progress of propolis for its biological and chemical compositions and its botanical origin. Evid. Based Complement. Alternat Med. 2013, 2013, 697390. [Google Scholar] [CrossRef] [PubMed]
  396. Huang, S.; Zhang, C.P.; Wang, K.; Li, G.Q.; Hu, F.L. Recent advances in the chemical composition of propolis. Molecules 2014, 19, 19610–19632. [Google Scholar] [CrossRef] [PubMed]
  397. Anjum, S.I.; Ullah, A.; Khan, K.A.; Attaullah, M.; Khan, H.; Ali, H.; Bashir, M.A.; Tahir, M.; Ansari, M.J.; Ghramh, H.A.; et al. Composition and functional properties of propolis (bee glue): A review. Saudi J. Biol. Sci. 2019, 26, 1695–1703. [Google Scholar] [CrossRef] [PubMed]
  398. Feng, L.; Cheah, I.K.; Ng, M.M.; Li, J.; Chan, S.M.; Lim, S.L.; Mahendran, R.; Kua, E.H.; Halliwell, B. The association between mushroom consumption and Mild Cognitive Impairment: A community-based cross-sectional study in Singapore. J. Alzheimers Dis. 2019, 68, 197–203. [Google Scholar] [CrossRef] [PubMed]
  399. Saral, Ö. An Investigation into Chestnut Honeys from Artvin Province in Turkiye: Their Physicochemical Properties, Phenolic Profiles and Antioxidant Activities. Chem. Biodivers. 2023, 20, e202201162. [Google Scholar] [CrossRef]
  400. Terzo, S.; Calvi, P.; Nuzzo, D.; Picone, P.; Galizzi, G.; Caruana, L.; Di Carlo, M.; Lentini, L.; Puleio, R.; Mule, F.; et al. Preventive Impact of Long-Term Ingestion of Chestnut Honey on Glucose Disorders and Neurodegeneration in Obese Mice. Nutrients 2022, 14, 756. [Google Scholar] [CrossRef] [PubMed]
  401. Terzo, S.; Calvi, P.; Nuzzo, D.; Picone, P.; Allegra, M.; Mule, F.; Amato, A. Long-Term Ingestion of Sicilian Black Bee Chestnut Honey and/or D-Limonene Counteracts Brain Damage Induced by High Fat-Diet in Obese Mice. Int. J. Mol. Sci. 2023, 24, 3467. [Google Scholar] [CrossRef] [PubMed]
  402. Sahin, H.; Kaltalioglu, K.; Erisgin, Z.; Coskun-Cevher, S.; Kolayli, S. Protective effects of aqueous extracts of some honeys against HCl/ethanol-induced gastric ulceration in rats. J. Food Biochem. 2019, 43, e13054. [Google Scholar] [CrossRef]
  403. Saral, Ö.; Yıldız, O.; Aliyazıcıoğlu, R.; Yuluğ, E.; Canpolat, S.; Öztürk, F.; Kolaylı, S. Apitherapy products enhance the recovery of CCL4-induced hepatic damages in rats. Turk. J. Med. Sci. 2016, 46, 194–202. [Google Scholar] [CrossRef]
  404. Seyhan, M.F.; Yılmaz, E.; Timirci-Kahraman, Ö.; Saygılı, N.; Kısakesen, H.I.; Eronat, A.P.; Ceviz, A.B.; Gazioğglu, S.B.; Yılmaz-Aydoğan, H.; Öztürk, O. Anatolian honey is not only sweet but can also protect from breast cancer: Elixir for women from Artemis to present. IUBMB Life 2017, 69, 677–688. [Google Scholar] [CrossRef] [PubMed]
  405. Kwon, E.B.; Kim, S.G.; Kim, Y.S.; Kim, B.; Han, S.M.; Lee, H.J.; Choi, H.M.; Choi, J.G. Castanea crenata honey reduces influenza infection by activating the innate immune response. Front. Immunol. 2023, 14, 1157506. [Google Scholar] [CrossRef]
  406. Oskay, G.S.; Oskay, D.; Ardac, N. Investigation of the Effect of Chestnut Honey and Curcumin Combination on Lifespan in the Experimental Heat Stress Model of Honey Bee. Biol. Bull. 2023, 50, 1393–1400. [Google Scholar] [CrossRef]
  407. Turska, M.; Pelak, J.; Turski, M.P.; Kocki, T.; Dukowski, P.; Plech, T.; Turski, W. Fate and distribution of kynurenic acid administered as beverage. Pharmacol. Rep. 2018, 70, 1089–1096. [Google Scholar] [CrossRef]
  408. Kaihara, M.; Price, J.M. The metabolism of quinaldic acid, kynurenic acid, and xanthurenic acid in the rabbit. J. Biol. Chem. 1962, 237, 1727–1729. [Google Scholar]
  409. Horibata, K.; Taniuchi, H.; Tashiro, M.; Kuno, S.; Hayaishi, O. The metabolism of kynurenic acid. II. Tracer experiments on the mechanism of kynurenic acid degradation and glutamic acid synthesis by Pseudomonas extracts. J. Biol. Chem. 1961, 236, 2991–2995. [Google Scholar]
  410. Kell, D.B.; Pretorius, E. No effects without causes. The Iron Dysregulation and Dormant Microbes hypothesis for chronic, inflammatory diseases. Biol. Rev. 2018, 93, 1518–1557. [Google Scholar] [PubMed]
  411. Butterfield, D.A.; Halliwell, B. Oxidative stress, dysfunctional glucose metabolism and Alzheimer disease. Nat. Rev. Neurosci. 2019, 20, 148–160. [Google Scholar] [CrossRef] [PubMed]
  412. Kell, D.B.; Heyden, E.L.; Pretorius, E. The biology of lactoferrin, an iron-binding protein that can help defend against viruses and bacteria. Frontiers Immunol. 2020, 11, 1221. [Google Scholar] [CrossRef]
  413. Kell, D.B.; Pretorius, E. The potential role of ischaemia-reperfusion injury in chronic, relapsing diseases such as rheumatoid arthritis, long COVID and ME/CFS: Evidence, mechanisms, and therapeutic implications. Biochem. J. 2022, 479, 1653–1708. [Google Scholar] [PubMed]
  414. Kell, D.B. Iron behaving badly: Inappropriate iron chelation as a major contributor to the aetiology of vascular and other progressive inflammatory and degenerative diseases. BMC Med. Genom. 2009, 2, 2. [Google Scholar]
  415. Kell, D.B. Towards a unifying, systems biology understanding of large-scale cellular death and destruction caused by poorly liganded iron: Parkinson’s, Huntington’s, Alzheimer’s, prions, bactericides, chemical toxicology and others as examples. Arch. Toxicol. 2010, 577, 825–889. [Google Scholar]
  416. Ramírez Ortega, D.; Ugalde Muñiz, P.E.; Blanco Ayala, T.; Vázquez Cervantes, G.I.; Lugo Huitrón, R.; Pineda, B.; Gonzalez Esquivel, D.F.; Pérez de la Cruz, G.; Pedraza Chaverri, J.; Sánchez Chapul, L.; et al. On the Antioxidant Properties of L-Kynurenine: An Efficient ROS Scavenger and Enhancer of Rat Brain Antioxidant Defense. Antioxidants 2021, 11, 31. [Google Scholar] [CrossRef] [PubMed]
  417. Worton, S.A.; Greenwood, S.L.; Wareing, M.; Heazell, A.E.; Myers, J. The kynurenine pathway; A new target for treating maternal features of preeclampsia? Placenta 2019, 84, 44–49. [Google Scholar] [CrossRef] [PubMed]
  418. Lugo-Huitrón, R.; Blanco-Ayala, T.; Ugalde-Muñiz, P.; Carrillo-Mora, P.; Pedraza-Chaverrí, J.; Silva-Adaya, D.; Maldonado, P.D.; Torres, I.; Pinzón, E.; Ortiz-Islas, E.; et al. On the antioxidant properties of kynurenic acid: Free radical scavenging activity and inhibition of oxidative stress. Neurotoxicol Teratol. 2011, 33, 538–547. [Google Scholar] [CrossRef] [PubMed]
  419. Reyes Ocampo, J.; Lugo Huitrón, R.; González-Esquivel, D.; Ugalde-Muñiz, P.; Jiménez-Anguiano, A.; Pineda, B.; Pedraza-Chaverri, J.; Ríos, C.; Pérez de la Cruz, V. Kynurenines with neuroactive and redox properties: Relevance to aging and brain diseases. Oxid. Med. Cell Longev. 2014, 2014, 646909. [Google Scholar] [CrossRef]
  420. Mor, A.; Tankiewicz-Kwedlo, A.; Krupa, A.; Pawlak, D. Role of Kynurenine Pathway in Oxidative Stress during Neurodegenerative Disorders. Cells 2021, 10, 1603. [Google Scholar] [CrossRef] [PubMed]
  421. Sheibani, M.; Shayan, M.; Khalilzadeh, M.; Soltani, Z.E.; Jafari-Sabet, M.; Ghasemi, M.; Dehpour, A.R. Kynurenine pathway and its role in neurologic, psychiatric, and inflammatory bowel diseases. Mol. Biol. Rep. 2023, 50, 10409–10425. [Google Scholar] [CrossRef] [PubMed]
  422. Darlington, L.G.; Mackay, G.M.; Forrest, C.M.; Stoy, N.; George, C.; Stone, T.W. Altered kynurenine metabolism correlates with infarct volume in stroke. Eur. J. Neurosci. 2007, 26, 2211–2221. [Google Scholar] [CrossRef] [PubMed]
  423. Majewski, M.; Kozlowska, A.; Thoene, M.; Lepiarczyk, E.; Grzegorzewski, W.J. Overview of the role of vitamins and minerals on the kynurenine pathway in health and disease. J. Physiol. Pharmacol. 2016, 67, 3–19. [Google Scholar]
  424. Sekine, A.; Fukuwatari, T. Acute liver failure increases kynurenic acid production in rat brain via changes in tryptophan metabolism in the periphery. Neurosci. Lett. 2019, 701, 14–19. [Google Scholar] [CrossRef]
  425. Hartai, Z.; Juhász, A.; Rimanóczy, Á.; Janáky, T.; Donkó, T.; Dux, L.; Penke, B.; Tóth, G.K.; Janka, Z.; Kálmán, J. Decreased serum and red blood cell kynurenic acid levels in Alzheimer’s disease. Neurochem. Int. 2007, 50, 308–313. [Google Scholar] [CrossRef]
  426. Heyes, M.P.; Saito, K.; Crowley, J.S.; Davis, L.E.; Demitrack, M.A.; Der, M.; Dilling, L.A.; Elia, J.; Kruesi, M.J.; Lackner, A.; et al. Quinolinic acid and kynurenine pathway metabolism in inflammatory and non-inflammatory neurological disease. Brain 1992, 115, 1249–1273. [Google Scholar] [CrossRef]
  427. Knapskog, A.B.; Aksnes, M.; Edwin, T.H.; Ueland, P.M.; Ulvik, A.; Fang, E.F.; Eldholm, R.S.; Halaas, N.B.; Saltvedt, I.; Giil, L.M.; et al. Higher concentrations of kynurenic acid in CSF are associated with the slower clinical progression of Alzheimer’s disease. Alzheimers Dement. 2023, 19, 5573–5582. [Google Scholar] [CrossRef]
  428. Liang, Y.; Xie, S.; He, Y.; Xu, M.; Qiao, X.; Zhu, Y.; Wu, W. Kynurenine Pathway Metabolites as Biomarkers in Alzheimer’s Disease. Dis. Markers 2022, 2022, 9484217. [Google Scholar] [CrossRef]
  429. Zapolski, T.; Kaminska, A.; Kocki, T.; Wysokinski, A.; Urbanska, E.M. Aortic stiffness-Is kynurenic acid a novel marker? Cross-sectional study in patients with persistent atrial fibrillation. PLoS ONE 2020, 15, e0236413. [Google Scholar] [CrossRef]
  430. Cavaleri, D.; Crocamo, C.; Morello, P.; Bartoli, F.; Carra, G. The Kynurenine Pathway in Attention-Deficit/Hyperactivity Disorder: A Systematic Review and Meta-Analysis of Blood Concentrations of Tryptophan and Its Catabolites. J. Clin. Med. 2024, 13, 583. [Google Scholar] [CrossRef]
  431. Liu, H.; Ding, L.; Zhang, H.; Mellor, D.; Wu, H.; Zhao, D.; Wu, C.; Lin, Z.; Yuan, J.; Peng, D. The Metabolic Factor Kynurenic Acid of Kynurenine Pathway Predicts Major Depressive Disorder. Front. Psychiatry 2018, 9, 552. [Google Scholar] [CrossRef]
  432. Colpo, G.D.; Venna, V.R.; McCullough, L.D.; Teixeira, A.L. Systematic Review on the Involvement of the Kynurenine Pathway in Stroke: Pre-clinical and Clinical Evidence. Front. Neurol. 2019, 10, 778. [Google Scholar] [CrossRef]
  433. Bartoli, F.; Cioni, R.M.; Cavaleri, D.; Callovini, T.; Crocamo, C.; Misiak, B.; Savitz, J.B.; Carra, G. The association of kynurenine pathway metabolites with symptom severity and clinical features of bipolar disorder: An overview. Eur. Psychiatry 2022, 65, e82. [Google Scholar] [CrossRef] [PubMed]
  434. Walczak, K.; Dabrowski, W.; Langner, E.; Zgrajka, W.; Piłat, J.; Kocki, T.; Rzeski, W.; Turski, W.A. Kynurenic acid synthesis and kynurenine aminotransferases expression in colon derived normal and cancer cells. Scand. J. Gastroenterol. 2011, 46, 903–912. [Google Scholar] [CrossRef]
  435. Walczak, K.; Turski, W.A.; Rajtar, G. Kynurenic acid inhibits colon cancer proliferation in vitro: Effects on signaling pathways. Amino Acids 2014, 46, 2393–2401. [Google Scholar] [CrossRef]
  436. Walczak, K.; Wnorowski, A.; Turski, W.A.; Plech, T. Kynurenic acid and cancer: Facts and controversies. Cell Mol. Life Sci. 2020, 77, 1531–1550. [Google Scholar] [CrossRef] [PubMed]
  437. Curto, M.; Lionetto, L.; Negro, A.; Capi, M.; Perugino, F.; Fazio, F.; Giamberardino, M.A.; Simmaco, M.; Nicoletti, F.; Martelletti, P. Altered serum levels of kynurenine metabolites in patients affected by cluster headache. J. Headache Pain 2015, 17, 27. [Google Scholar] [CrossRef]
  438. Curto, M.; Lionetto, L.; Negro, A.; Capi, M.; Fazio, F.; Giamberardino, M.A.; Simmaco, M.; Nicoletti, F.; Martelletti, P. Altered kynurenine pathway metabolites in serum of chronic migraine patients. J. Headache Pain 2015, 17, 47. [Google Scholar] [CrossRef]
  439. Badawy, A.A. The kynurenine pathway of tryptophan metabolism: A neglected therapeutic target of COVID-19 pathophysiology and immunotherapy. Biosci. Rep. 2023, 43, BSR20230595. [Google Scholar] [CrossRef]
  440. Costanzo, M.; Caterino, M.; Fedele, R.; Cevenini, A.; Pontillo, M.; Barra, L.; Ruoppolo, M. COVIDomics: The Proteomic and Metabolomic Signatures of COVID-19. Int. J. Mol. Sci. 2022, 23, 2414. [Google Scholar] [CrossRef]
  441. Tezcan, D.; Onmaz, D.E.; Sivrikaya, A.; Körez, M.K.; Hakbilen, S.; Gülcemal, S.; Yılmaz, S. Kynurenine pathway of tryptophan metabolism in patients with familial Mediterranean fever. Mod. Rheumatol. 2023, 33, 398–407. [Google Scholar] [CrossRef] [PubMed]
  442. Al Saedi, A.; Chow, S.; Vogrin, S.; Guillemin, G.J.; Duque, G. Association Between Tryptophan Metabolites, Physical Performance, and Frailty in Older Persons. Int. J. Tryptophan Res. 2022, 15, 11786469211069951. [Google Scholar] [CrossRef] [PubMed]
  443. Beal, M.F.; Matson, W.R.; Storey, E.; Milbury, P.; Ryan, E.A.; Ogawa, T.; Bird, E.D. Kynurenic acid concentrations are reduced in Huntington’s disease cerebral cortex. J. Neurol. Sci. 1992, 108, 80–87. [Google Scholar] [CrossRef] [PubMed]
  444. Kaye, J.; Piryatinsky, V.; Birnberg, T.; Hingaly, T.; Raymond, E.; Kashi, R.; Amit-Romach, E.; Caballero, I.S.; Towfic, F.; Ator, M.A.; et al. Laquinimod arrests experimental autoimmune encephalomyelitis by activating the aryl hydrocarbon receptor. Proc. Natl. Acad. Sci. USA 2016, 113, E6145–E6152. [Google Scholar] [CrossRef]
  445. Shen, H.; Xu, X.; Bai, Y.; Wang, X.; Wu, Y.; Zhong, J.; Wu, Q.; Luo, Y.; Shang, T.; Shen, R.; et al. Therapeutic potential of targeting kynurenine pathway in neurodegenerative diseases. Eur. J. Med. Chem. 2023, 251, 115258. [Google Scholar] [CrossRef]
  446. Tsuji, A.; Ikeda, Y.; Yoshikawa, S.; Taniguchi, K.; Sawamura, H.; Morikawa, S.; Nakashima, M.; Asai, T.; Matsuda, S. The Tryptophan and Kynurenine Pathway Involved in the Development of Immune-Related Diseases. Int. J. Mol. Sci. 2023, 24, 5742. [Google Scholar] [CrossRef]
  447. Li, P.; Zheng, J.; Bai, Y.; Wang, D.; Cui, Z.; Li, Y.; Zhang, J.; Wang, Y. Characterization of kynurenine pathway in patients with diarrhea-predominant irritable bowel syndrome. Eur. J. Histochem. 2020, 64, 3132. [Google Scholar] [CrossRef]
  448. Chojnacki, C.; Błońska, A.; Konrad, P.; Chojnacki, M.; Podogrocki, M.; Poplawski, T. Changes in Tryptophan Metabolism on Serotonin and Kynurenine Pathways in Patients with Irritable Bowel Syndrome. Nutrients 2023, 15, 1262. [Google Scholar] [CrossRef]
  449. Pawlak, D.; Pawlak, K.; Malyszko, J.; Mysliwiec, M.; Buczko, W. Accumulation of toxic products degradation of kynurenine in hemodialyzed patients. Int. Urol. Nephrol. 2001, 33, 399–404. [Google Scholar] [CrossRef]
  450. Krupa, A.; Krupa, M.M.; Pawlak, K. Kynurenine Pathway-An Underestimated Factor Modulating Innate Immunity in Sepsis-Induced Acute Kidney Injury? Cells 2022, 11, 2604. [Google Scholar] [CrossRef]
  451. Mor, A.; Kalaska, B.; Pawlak, D. Kynurenine Pathway in Chronic Kidney Disease: What’s Old, What’s New, and What’s Next? Int. J. Tryptophan Res. 2020, 13, 1178646920954882. [Google Scholar] [CrossRef] [PubMed]
  452. Zhao, J. Plasma kynurenic acid/tryptophan ratio: A sensitive and reliable biomarker for the assessment of renal function. Ren. Fail. 2013, 35, 648–653. [Google Scholar] [CrossRef]
  453. Erabi, H.; Okada, G.; Shibasaki, C.; Setoyama, D.; Kang, D.; Takamura, M.; Yoshino, A.; Fuchikami, M.; Kurata, A.; Kato, T.A.; et al. Kynurenic acid is a potential overlapped biomarker between diagnosis and treatment response for depression from metabolome analysis. Sci. Rep. 2020, 10, 16822. [Google Scholar] [CrossRef]
  454. Tanaka, M.; Bohár, Z.; Martos, D.; Telegdy, G.; Vécsei, L. Antidepressant-like effects of kynurenic acid in a modified forced swim test. Pharmacol. Rep. 2020, 72, 449–455. [Google Scholar] [CrossRef]
  455. Francis, H.M.; Stevenson, R.J.; Tan, L.S.Y.; Ehrenfeld, L.; Byeon, S.; Attuquayefio, T.; Gupta, D.; Lim, C.K. Kynurenic acid as a biochemical factor underlying the association between Western-style diet and depression: A cross-sectional study. Front. Nutr. 2022, 9, 945538. [Google Scholar] [CrossRef]
  456. Li, C.C.; Ye, F.; Xu, C.X.; Jiang, N.; Chang, Q.; Liu, X.M.; Pan, R.L. Tryptophan-kynurenine metabolic characterization in the gut and brain of depressive-like rats induced by chronic restraint stress. J. Affect. Disord. 2023, 328, 273–286. [Google Scholar] [CrossRef]
  457. Fila, M.; Chojnacki, J.; Derwich, M.; Chojnacki, C.; Pawlowska, E.; Blasiak, J. Urine 5-Hydroxyindoleacetic Acid Negatively Correlates with Migraine Occurrence and Characteristics in the Interictal Phase of Episodic Migraine. Int. J. Mol. Sci. 2024, 25, 5471. [Google Scholar] [CrossRef]
  458. Biringer, R.G. Migraine signaling pathways: Amino acid metabolites that regulate migraine and predispose migraineurs to headache. Mol. Cell Biochem. 2022, 477, 2269–2296. [Google Scholar] [CrossRef] [PubMed]
  459. Hartai, Z.; Klivényi, P.; Janáky, T.; Penke, B.; Dux, L.; Vécsei, L. Kynurenine metabolism in multiple sclerosis. Acta Neurol. Scand. 2005, 112, 93–96. [Google Scholar] [CrossRef]
  460. Lim, C.K.; Bilgin, A.; Lovejoy, D.B.; Tan, V.; Bustamante, S.; Taylor, B.V.; Bessede, A.; Brew, B.J.; Guillemin, G.J. Kynurenine pathway metabolomics predicts and provides mechanistic insight into multiple sclerosis progression. Sci. Rep. 2017, 7, 41473. [Google Scholar] [CrossRef]
  461. Sandi, D.; Fricska-Nagy, Z.; Bencsik, K.; Vécsei, L. Neurodegeneration in Multiple Sclerosis: Symptoms of Silent Progression, Biomarkers and Neuroprotective Therapy-Kynurenines Are Important Players. Molecules 2021, 26, 3423. [Google Scholar] [CrossRef] [PubMed]
  462. Rejdak, K.; Bartosik-Psujek, H.; Dobosz, B.; Kocki, T.; Grieb, P.; Giovannoni, G.; Turski, W.A.; Stelmasiak, Z. Decreased level of kynurenic acid in cerebrospinal fluid of relapsing-onset multiple sclerosis patients. Neurosci. Lett. 2002, 331, 63–65. [Google Scholar] [CrossRef]
  463. Dehhaghi, M.; Panahi, H.K.S.; Kavyani, B.; Heng, B.; Tan, V.; Braidy, N.; Guillemin, G.J. The Role of Kynurenine Pathway and NAD(+) Metabolism in Myalgic Encephalomyelitis/Chronic Fatigue Syndrome. Aging Dis. 2022, 13, 698–711. [Google Scholar] [CrossRef]
  464. Kavyani, B.; Lidbury, B.A.; Schloeffel, R.; Fisher, P.R.; Missailidis, D.; Annesley, S.J.; Dehhaghi, M.; Heng, B.; Guillemin, G.J. Could the kynurenine pathway be the key missing piece of Myalgic Encephalomyelitis/Chronic Fatigue Syndrome (ME/CFS) complex puzzle? Cell Mol. Life Sci. 2022, 79, 412. [Google Scholar] [CrossRef] [PubMed]
  465. Kavyani, B.; Ahn, S.B.; Missailidis, D.; Annesley, S.J.; Fisher, P.R.; Schloeffel, R.; Guillemin, G.J.; Lovejoy, D.B.; Heng, B. Dysregulation of the Kynurenine Pathway, Cytokine Expression Pattern, and Proteomics Profile Link to Symptomology in Myalgic Encephalomyelitis/Chronic Fatigue Syndrome (ME/CFS). Mol. Neurobiol. 2024, 61, 3771–3787. [Google Scholar] [CrossRef]
  466. Shi, T.; Shi, Y.; Gao, H.; Ma, Y.; Wang, Q.; Shen, S.; Shao, X.; Gong, W.; Chen, X.; Qin, J.; et al. Exercised accelerated the production of muscle-derived kynurenic acid in skeletal muscle and alleviated the postmenopausal osteoporosis through the Gpr35/NFκB p65 pathway. J. Orthop. Transl. 2022, 35, 1–12. [Google Scholar] [CrossRef]
  467. Lim, C.K.; Fernández-Gomez, F.J.; Braidy, N.; Estrada, C.; Costa, C.; Costa, S.; Bessede, A.; Fernandez-Villalba, E.; Zinger, A.; Herrero, M.T.; et al. Involvement of the kynurenine pathway in the pathogenesis of Parkinson’s disease. Prog. Neurobiol. 2017, 155, 76–95. [Google Scholar] [CrossRef]
  468. Ogawa, T.; Matson, W.R.; Beal, M.F.; Myers, R.H.; Bird, E.D.; Milbury, P.; Saso, S. Kynurenine pathway abnormalities in Parkinson’s disease. Neurology 1992, 42, 1702–1706. [Google Scholar] [CrossRef]
  469. Pires, A.S.; Gupta, S.; Barton, S.A.; Vander Wall, R.; Tan, V.; Heng, B.; Phillips, J.K.; Guillemin, G.J. Temporal Profile of Kynurenine Pathway Metabolites in a Rodent Model of Autosomal Recessive Polycystic Kidney Disease. Int. J. Tryptophan Res. 2022, 15, 11786469221126063. [Google Scholar] [CrossRef] [PubMed]
  470. Wang, S.; Mu, L.; Zhang, C.; Long, X.; Zhang, Y.; Li, R.; Zhao, Y.; Qiao, J. Abnormal Activation of Tryptophan-Kynurenine Pathway in Women With Polycystic Ovary Syndrome. Front. Endocrinol. 2022, 13, 877807. [Google Scholar] [CrossRef]
  471. Van Zundert, S.K.; Broekhuizen, M.; Smit, A.J.; van Rossem, L.; Mirzaian, M.; Willemsen, S.P.; Danser, A.J.; De Rijke, Y.B.; Reiss, I.K.; Merkus, D.; et al. The Role of the Kynurenine Pathway in the (Patho) physiology of Maternal Pregnancy and Fetal Outcomes: A Systematic Review. Int. J. Tryptophan Res. 2022, 15, 11786469221135545. [Google Scholar] [CrossRef]
  472. Cai, Z.; Tian, S.; Klein, T.; Tu, L.; Geenen, L.W.; Koudstaal, T.; van den Bosch, A.E.; de Rijke, Y.B.; Reiss, I.K.M.; Boersma, E.; et al. Kynurenine metabolites predict survival in pulmonary arterial hypertension: A role for IL-6/IL-6Ralpha. Sci. Rep. 2022, 12, 12326. [Google Scholar] [CrossRef]
  473. Linderholm, K.R.; Skogh, E.; Olsson, S.K.; Dahl, M.L.; Holtze, M.; Engberg, G.; Samuelsson, M.; Erhardt, S. Increased levels of kynurenine and kynurenic acid in the CSF of patients with schizophrenia. Schizophr. Bull. 2012, 38, 426–432. [Google Scholar] [CrossRef] [PubMed]
  474. Erhardt, S.; Schwieler, L.; Imbeault, S.; Engberg, G. The kynurenine pathway in schizophrenia and bipolar disorder. Neuropharmacology 2017, 112, 297–306. [Google Scholar] [CrossRef]
  475. Kozak, R.; Campbell, B.M.; Strick, C.A.; Horner, W.; Hoffmann, W.E.; Kiss, T.; Chapin, D.S.; McGinnis, D.; Abbott, A.L.; Roberts, B.M.; et al. Reduction of brain kynurenic acid improves cognitive function. J. Neurosci. 2014, 34, 10592–10602. [Google Scholar] [CrossRef] [PubMed]
  476. Tanaka, M.; Tóth, F.; Polyák, H.; Szabó, Á.; Mándi, Y.; Vécsei, L. Immune Influencers in Action: Metabolites and Enzymes of the Tryptophan-Kynurenine Metabolic Pathway. Biomedicines 2021, 9, 734. [Google Scholar] [CrossRef] [PubMed]
  477. Kegel, M.E.; Johansson, V.; Wetterberg, L.; Bhat, M.; Schwieler, L.; Cannon, T.D.; Schuppe-Koistinen, I.; Engberg, G.; Landén, M.; Hultman, C.M.; et al. Kynurenic acid and psychotic symptoms and personality traits in twins with psychiatric morbidity. Psychiatry Res. 2017, 247, 105–112. [Google Scholar] [CrossRef]
  478. Chen, W.; Tian, Y.; Gou, M.; Wang, L.; Tong, J.; Zhou, Y.; Feng, W.; Li, Y.; Chen, S.; Liu, Y.; et al. Role of the immune-kynurenine pathway in treatment-resistant schizophrenia. Prog. Neuropsychopharmacol. Biol. Psychiatry 2024, 130, 110926. [Google Scholar] [CrossRef]
  479. Skorobogatov, K.; Autier, V.; Foiselle, M.; Richard, J.R.; Boukouaci, W.; Wu, C.L.; Raynal, S.; Carbonne, C.; Laukens, K.; Meysman, P.; et al. Kynurenine pathway abnormalities are state-specific but not diagnosis-specific in schizophrenia and bipolar disorder. Brain Behav. Immun. Health 2023, 27, 100584. [Google Scholar] [CrossRef]
  480. Eryavuz Onmaz, D.; Tezcan, D.; Abusoglu, S.; Sak, F.; Humeyra Yerlikaya, F.; Yilmaz, S.; Abusoglu, G.; Kazim Korez, M.; Unlu, A. Impaired kynurenine metabolism in patients with primary Sjögren’s syndrome. Clin. Biochem. 2023, 114, 1–10. [Google Scholar] [CrossRef] [PubMed]
  481. Eryavuz Onmaz, D.; Tezcan, D.; Yilmaz, S.; Onmaz, M.; Unlu, A. Altered kynurenine pathway metabolism and association with disease activity in patients with systemic lupus. Amino Acids 2023, 55, 1937–1947. [Google Scholar] [CrossRef] [PubMed]
  482. Chmiel-Perzyńska, I.; Perzyński, A.; Urbańska, E.M. Experimental diabetes mellitus type 1 increases hippocampal content of kynurenic acid in rats. Pharmacol. Rep. 2014, 66, 1134–1139. [Google Scholar] [CrossRef]
  483. Oxenkrug, G.F. Increased Plasma Levels of Xanthurenic and Kynurenic Acids in Type 2 Diabetes. Mol. Neurobiol. 2015, 52, 805–810. [Google Scholar] [CrossRef] [PubMed]
  484. Kiluk, M.; Lewkowicz, J.; Pawlak, D.; Tankiewicz-Kwedlo, A. Crosstalk between Tryptophan Metabolism via Kynurenine Pathway and Carbohydrate Metabolism in the Context of Cardio-Metabolic Risk-Review. J. Clin. Med. 2021, 10, 2484. [Google Scholar] [CrossRef]
  485. Sofia, M.A.; Ciorba, M.A.; Meckel, K.; Lim, C.K.; Guillemin, G.J.; Weber, C.R.; Bissonnette, M.; Pekow, J.R. Tryptophan Metabolism through the Kynurenine Pathway is Associated with Endoscopic Inflammation in Ulcerative Colitis. Inflamm. Bowel Dis. 2018, 24, 1471–1480. [Google Scholar] [CrossRef]
  486. Davis, H.E.; McCorkell, L.; Vogel, J.M.; Topol, E.J. Long COVID: Major findings, mechanisms and recommendations. Nat. Rev. Microbiol. 2023, 21, 133–146. [Google Scholar] [CrossRef]
  487. Proal, A.D.; VanElzakker, M.B. Long COVID or Post-acute Sequelae of COVID-19 (PASC): An Overview of Biological Factors That May Contribute to Persistent Symptoms. Front. Microbiol. 2021, 12, 698169. [Google Scholar] [CrossRef]
  488. Thomas, T.; Stefanoni, D.; Reisz, J.A.; Nemkov, T.; Bertolone, L.; Francis, R.O.; Hudson, K.E.; Zimring, J.C.; Hansen, K.C.; Hod, E.A.; et al. COVID-19 infection alters kynurenine and fatty acid metabolism, correlating with IL-6 levels and renal status. JCI Insight 2020, 5, e140327. [Google Scholar] [CrossRef]
  489. Roberts, I.; Wright Muelas, M.; Taylor, J.M.; Davison, A.S.; Xu, Y.; Grixti, J.M.; Gotts, N.; Sorokin, A.; Goodacre, R.; Kell, D.B. Untargeted metabolomics of COVID-19 patient serum reveals potential prognostic markers of both severity and outcome. Metabolomics 2022, 18, 6. [Google Scholar] [CrossRef]
  490. Cihan, M.; Doğan, Ö.; Ceran Serdar, C.; Altunçekiç Yıldırım, A.; Kurt, C.; Serdar, M.A. Kynurenine pathway in Coronavirus disease (COVID-19): Potential role in prognosis. J. Clin. Lab. Anal. 2022, 36, e24257. [Google Scholar] [CrossRef]
  491. Kucukkarapinar, M.; Yay-Pence, A.; Yildiz, Y.; Buyukkoruk, M.; Yaz-Aydin, G.; Deveci-Bulut, T.S.; Gulbahar, O.; Senol, E.; Candansayar, S. Psychological outcomes of COVID-19 survivors at sixth months after diagnose: The role of kynurenine pathway metabolites in depression, anxiety, and stress. J. Neural Transm. 2022, 129, 1077–1089. [Google Scholar] [CrossRef]
  492. Sindelar, M.; Stancliffe, E.; Schwaiger-Haber, M.; Anbukumar, D.S.; Adkins-Travis, K.; Goss, C.W.; O’Halloran, J.A.; Mudd, P.A.; Liu, W.C.; Albrecht, R.A.; et al. Longitudinal metabolomics of human plasma reveals prognostic markers of COVID-19 disease severity. Cell Rep. Med. 2021, 2, 100369. [Google Scholar] [CrossRef]
  493. Cai, Y.; Kim, D.J.; Takahashi, T.; Broadhurst, D.I.; Yan, H.; Ma, S.; Rattray, N.J.W.; Casanovas-Massana, A.; Israelow, B.; Klein, J.; et al. Kynurenic acid may underlie sex-specific immune responses to COVID-19. Sci. Signal 2021, 14, eabf8483. [Google Scholar] [CrossRef] [PubMed]
  494. Almulla, A.F.; Supasitthumrong, T.; Tunvirachaisakul, C.; Algon, A.A.A.; Al-Hakeim, H.K.; Maes, M. The tryptophan catabolite or kynurenine pathway in COVID-19 and critical COVID-19: A systematic review and meta-analysis. BMC Infect. Dis. 2022, 22, 615. [Google Scholar] [CrossRef]
  495. Holmes, E.; Wist, J.; Masuda, R.; Lodge, S.; Nitschke, P.; Kimhofer, T.; Loo, R.L.; Begum, S.; Boughton, B.; Yang, R.; et al. Incomplete Systemic Recovery and Metabolic Phenoreversion in Post-Acute-Phase Nonhospitalized COVID-19 Patients: Implications for Assessment of Post-Acute COVID-19 Syndrome. J. Proteome Res. 2021, 20, 3315–3329. [Google Scholar] [CrossRef]
  496. Overmyer, K.A.; Shishkova, E.; Miller, I.J.; Balnis, J.; Bernstein, M.N.; Peters-Clarke, T.M.; Meyer, J.G.; Quan, Q.; Muehlbauer, L.K.; Trujillo, E.A.; et al. Large-Scale Multi-omic Analysis of COVID-19 Severity. Cell Syst. 2020, 12, 23–40e27. [Google Scholar] [CrossRef]
  497. Bizjak, D.A.; Stangl, M.; Börner, N.; Bösch, F.; Durner, J.; Drunin, G.; Buhl, J.L.; Abendroth, D. Kynurenine serves as useful biomarker in acute, Long- and Post-COVID-19 diagnostics. Front. Immunol. 2022, 13, 1004545. [Google Scholar] [CrossRef]
  498. Mangge, H.; Herrmann, M.; Meinitzer, A.; Pailer, S.; Curcic, P.; Sloup, Z.; Holter, M.; Prüller, F. Increased Kynurenine Indicates a Fatal Course of COVID-19. Antioxidants 2021, 10, 1960. [Google Scholar] [CrossRef]
  499. Michaelis, S.; Zelzer, S.; Schnedl, W.J.; Baranyi, A.; Meinitzer, A.; Enko, D. Assessment of tryptophan and kynurenine as prognostic markers in patients with SARS-CoV-2. Clin. Chim. Acta 2022, 525, 29–33. [Google Scholar] [CrossRef] [PubMed]
  500. Michaelis, S.; Zelzer, S.; Schneider, C.; Schnedl, W.J.; Baranyi, A.; Meinitzer, A.; Herrmann, M.; Enko, D. Alteration of the kynurenine pathway is inversely associated with the humoral immune response in patients with SARS-CoV-2. Clin. Chim. Acta 2022, 537, 77–79. [Google Scholar] [CrossRef] [PubMed]
  501. Ceballos, F.C.; Virseda-Berdices, A.; Resino, S.; Ryan, P.; Martinez-Gonzalez, O.; Perez-Garcia, F.; Martin-Vicente, M.; Brochado-Kith, O.; Blancas, R.; Bartolome-Sanchez, S.; et al. Metabolic Profiling at COVID-19 Onset Shows Disease Severity and Sex-Specific Dysregulation. Front. Immunol. 2022, 13, 925558. [Google Scholar] [CrossRef]
  502. Dewulf, J.P.; Martin, M.; Marie, S.; Oguz, F.; Belkhir, L.; De Greef, J.; Yombi, J.C.; Wittebole, X.; Laterre, P.F.; Jadoul, M.; et al. Urine metabolomics links dysregulation of the tryptophan-kynurenine pathway to inflammation and severity of COVID-19. Sci. Rep. 2022, 12, 9959. [Google Scholar] [CrossRef]
  503. Brown, M.; Dunn, W.B.; Ellis, D.I.; Goodacre, R.; Handl, J.; Knowles, J.D.; O’Hagan, S.; Spasic, I.; Kell, D.B. A metabolome pipeline: From concept to data to knowledge. Metabolomics 2005, 1, 39–51. [Google Scholar]
  504. Kell, D.B.; Oliver, S.G. Here is the evidence, now what is the hypothesis? The complementary roles of inductive and hypothesis-driven science in the post-genomic era. Bioessays 2004, 26, 99–105. [Google Scholar] [PubMed]
  505. Li, R.W.S.; Yang, C.; Sit, A.S.M.; Kwan, Y.W.; Lee, S.M.Y.; Hoi, M.P.M.; Chan, S.W.; Hausman, M.; Vanhoutte, P.M.; Leung, G.P.H. Uptake and Protective Effects of Ergothioneine in Human Endothelial Cells. J. Pharmacol. Exp. Ther. 2014, 350, 691–700. [Google Scholar] [CrossRef] [PubMed]
  506. Paul, B.D.; Snyder, S.H. The unusual amino acid L-ergothioneine is a physiologic cytoprotectant. Cell Death Differ. 2010, 17, 1134–1140. [Google Scholar] [CrossRef] [PubMed]
  507. Cheah, I.K.; Halliwell, B. Ergothioneine; antioxidant potential, physiological function and role in disease. Biochim. Biophys. Acta 2012, 1822, 784–793. [Google Scholar] [CrossRef] [PubMed]
  508. Halliwell, B.; Cheah, I.K.; Tang, R.M.Y. Ergothioneine—A diet-derived antioxidant with therapeutic potential. FEBS Lett. 2018, 592, 3357–3366. [Google Scholar] [CrossRef] [PubMed]
  509. Connick, J.H.; Heywood, G.C.; Sills, G.J.; Thompson, G.G.; Brodie, M.J.; Stone, T.W. Nicotinylalanine increases cerebral kynurenic acid content and has anticonvulsant activity. Gen. Pharmacol. 1992, 23, 235–239. [Google Scholar] [CrossRef] [PubMed]
  510. Stone, T.W. Development and therapeutic potential of kynurenic acid and kynurenine derivatives for neuroprotection. Trends Pharmacol. Sci. 2000, 21, 149–154. [Google Scholar] [CrossRef] [PubMed]
  511. Kocki, T.; Wielosz, M.; Turski, W.A.; Urbanska, E.M. Enhancement of brain kynurenic acid production by anticonvulsants--novel mechanism of antiepileptic activity? Eur. J. Pharmacol. 2006, 541, 147–151. [Google Scholar] [CrossRef] [PubMed]
  512. Brown, S.J.; Huang, X.F.; Newell, K.A. The kynurenine pathway in major depression: What we know and where to next. Neurosci. Biobehav. Rev. 2021, 127, 917–927. [Google Scholar] [CrossRef]
  513. Yan, J.; Kothur, K.; Innes, E.A.; Han, V.X.; Jones, H.F.; Patel, S.; Tsang, E.; Webster, R.; Gupta, S.; Troedson, C.; et al. Decreased cerebrospinal fluid kynurenic acid in epileptic spasms: A biomarker of response to corticosteroids. EBioMedicine 2022, 84, 104280. [Google Scholar] [CrossRef]
  514. Thaler, S.; Choragiewicz, T.J.; Rejdak, R.; Fiedorowicz, M.; Turski, W.A.; Tulidowicz-Bielak, M.; Zrenner, E.; Schuettauf, F.; Zarnowski, T. Neuroprotection by acetoacetate and beta-hydroxybutyrate against NMDA-induced RGC damage in rat--possible involvement of kynurenic acid. Graefes Arch. Clin. Exp. Ophthalmol. 2010, 248, 1729–1735. [Google Scholar] [CrossRef]
  515. Chmiel-Perzyńska, I.; Kloc, R.; Perzyński, A.; Rudzki, S.; Urbańska, E.M. Novel aspect of ketone action: Beta-hydroxybutyrate increases brain synthesis of kynurenic acid in vitro. Neurotox. Res. 2011, 20, 40–50. [Google Scholar] [CrossRef]
  516. Salvati, P.; Ukmar, G.; Dho, L.; Rosa, B.; Cini, M.; Marconi, M.; Molinari, A.; Post, C. Brain concentrations of kynurenic acid after a systemic neuroprotective dose in the gerbil model of global ischemia. Prog. Neuropsychopharmacol. Biol. Psychiatry 1999, 23, 741–752. [Google Scholar] [CrossRef]
  517. Andiné, P.; Lehmann, A.; Ellren, K.; Wennberg, E.; Kjellmer, I.; Nielsen, T.; Hagberg, H. The excitatory amino acid antagonist kynurenic acid administered after hypoxic-ischemia in neonatal rats offers neuroprotection. Neurosci. Lett. 1988, 90, 208–212. [Google Scholar] [CrossRef]
  518. Bratek-Gerej, E.; Ziembowicz, A.; Godlewski, J.; Salinska, E. The Mechanism of the Neuroprotective Effect of Kynurenic Acid in the Experimental Model of Neonatal Hypoxia-Ischemia: The Link to Oxidative Stress. Antioxidants 2021, 10, 1775. [Google Scholar] [CrossRef]
  519. Gigler, G.; Szénási, G.; Simó, A.; Lévay, G.; Hársing, L.G., Jr.; Sas, K.; Vécsei, L.; Toldi, J. Neuroprotective effect of L-kynurenine sulfate administered before focal cerebral ischemia in mice and global cerebral ischemia in gerbils. Eur. J. Pharmacol. 2007, 564, 116–122. [Google Scholar] [CrossRef] [PubMed]
  520. Robotka, H.; Sas, K.; Ágoston, M.; Rózsa, É.; Szénási, G.; Gigler, G.; Vécsei, L.; Toldi, J. Neuroprotection achieved in the ischaemic rat cortex with L-kynurenine sulphate. Life Sci. 2008, 82, 915–919. [Google Scholar] [CrossRef]
  521. Martos, D.; Tuka, B.; Tanaka, M.; Vécsei, L.; Telegdy, G. Memory Enhancement with Kynurenic Acid and Its Mechanisms in Neurotransmission. Biomedicines 2022, 10, 849. [Google Scholar] [CrossRef]
  522. Vohra, M.; Lemieux, G.A.; Lin, L.; Ashrafi, K. Kynurenic acid accumulation underlies learning and memory impairment associated with aging. Genes Dev. 2018, 32, 14–19. [Google Scholar] [CrossRef]
  523. Fejes, A.; Párdutz, Á.; Toldi, J.; Vécsei, L. Kynurenine metabolites and migraine: Experimental studies and therapeutic perspectives. Curr. Neuropharmacol. 2011, 9, 376–387. [Google Scholar] [CrossRef]
  524. Párdutz, Á.; Fejes, A.; Bohár, Z.; Tar, L.; Toldi, J.; Vécsei, L. Kynurenines and headache. J. Neural Transm. 2012, 119, 285–296. [Google Scholar] [CrossRef]
  525. Oláh, G.; Heredi, J.; Menyhart, A.; Czinege, Z.; Nagy, D.; Fuzik, J.; Kocsis, K.; Knapp, L.; Krucsó, E.; Géllert, L.; et al. Unexpected effects of peripherally administered kynurenic acid on cortical spreading depression and related blood-brain barrier permeability. Drug Des. Devel Ther. 2013, 7, 981–987. [Google Scholar] [CrossRef]
  526. Körtési, T.; Spekker, E.; Vécsei, L. Exploring the Tryptophan Metabolic Pathways in Migraine-Related Mechanisms. Cells 2022, 11, 3795. [Google Scholar] [CrossRef] [PubMed]
  527. Lovelace, M.D.; Varney, B.; Sundaram, G.; Franco, N.F.; Ng, M.L.; Pai, S.; Lim, C.K.; Guillemin, G.J.; Brew, B.J. Current Evidence for a Role of the Kynurenine Pathway of Tryptophan Metabolism in Multiple Sclerosis. Front. Immunol. 2016, 7, 246. [Google Scholar] [CrossRef]
  528. Lovelace, M.D.; Varney, B.; Sundaram, G.; Lennon, M.J.; Lim, C.K.; Jacobs, K.; Guillemin, G.J.; Brew, B.J. Recent evidence for an expanded role of the kynurenine pathway of tryptophan metabolism in neurological diseases. Neuropharmacology 2017, 112, 373–388. [Google Scholar] [CrossRef]
  529. Rojewska, E.; Ciapała, K.; Mika, J. Kynurenic acid and zaprinast diminished CXCL17-evoked pain-related behaviour and enhanced morphine analgesia in a mouse neuropathic pain model. Pharmacol. Rep. 2019, 71, 139–148. [Google Scholar] [CrossRef] [PubMed]
  530. Samavati, R.; Zádor, F.; Szűcs, E.; Tuka, B.; Martos, D.; Veres, G.; Gáspár, R.; Mándity, I.M.; Fülöp, F.; Vécsei, L.; et al. Kynurenic acid and its analogue can alter the opioid receptor G-protein signaling after acute treatment via NMDA receptor in rat cortex and striatum. J. Neurol. Sci. 2017, 376, 63–70. [Google Scholar] [CrossRef] [PubMed]
  531. Ciapała, K.; Mika, J.; Rojewska, E. The Kynurenine Pathway as a Potential Target for Neuropathic Pain Therapy Design: From Basic Research to Clinical Perspectives. Int. J. Mol. Sci. 2021, 22, 11055. [Google Scholar] [CrossRef] [PubMed]
  532. Tanaka, M.; Bohár, Z.; Vécsei, L. Are Kynurenines Accomplices or Principal Villains in Dementia? Maintenance of Kynurenine Metabolism. Molecules 2020, 25, 564. [Google Scholar] [CrossRef] [PubMed]
  533. Ostapiuk, A.; Urbanska, E.M. Kynurenic acid in neurodegenerative disorders-unique neuroprotection or double-edged sword? CNS Neurosci. Ther. 2022, 28, 19–35. [Google Scholar] [CrossRef] [PubMed]
  534. Korimová, A.; Cížková, D.; Toldi, J.; Vécsei, L.; Vanický, I. Protective effects of glucosamine-kynurenic acid after compression-induced spinal cord injury in the rat. Centr Eur. J. Biol. 2012, 7, 996–1004. [Google Scholar] [CrossRef]
  535. Alme, K.N.; Ulvik, A.; Askim, T.; Assmus, J.; Mollnes, T.E.; Naik, M.; Naess, H.; Saltvedt, I.; Ueland, P.M.; Knapskog, A.B. Neopterin and kynurenic acid as predictors of stroke recurrence and mortality: A multicentre prospective cohort study on biomarkers of inflammation measured three months after ischemic stroke. BMC Neurol. 2021, 21, 476. [Google Scholar] [CrossRef]
  536. Vieira, J.P.P.; Karampatsi, D.; Vercalsteren, E.; Darsalia, V.; Patrone, C.; Duarte, J.M.N. Nuclear magnetic resonance spectroscopy reveals biomarkers of stroke recovery in a mouse model of obesity-associated type 2 diabetes. Biosci. Rep. 2024, 44, BSR20240249. [Google Scholar] [CrossRef]
  537. Mangas, A.; Heredia, M.; Riolobos, A.; De la Fuente, A.; Criado, J.M.; Yajeya, J.; Geffard, M.; Coveñas, R. Overexpression of kynurenic acid and 3-hydroxyanthranilic acid after rat traumatic brain injury. Eur. J. Histochem. 2018, 62, 278–284. [Google Scholar] [CrossRef]
  538. Suma, T.; Koshinaga, M.; Fukushima, M.; Kano, T.; Katayama, Y. Effects of in situ administration of excitatory amino acid antagonists on rapid microglial and astroglial reactions in rat hippocampus following traumatic brain injury. Neurol. Res. 2008, 30, 420–429. [Google Scholar] [CrossRef]
  539. Kell, D.B.; Laubscher, G.J.; Pretorius, E. A central role for amyloid fibrin microclots in long COVID/PASC: Origins and therapeutic implications. Biochem. J. 2022, 479, 537–559. [Google Scholar] [CrossRef]
  540. Pretorius, E.; Kell, D.B. A perspective on how microscopy imaging of fibrinaloid microclots and platelet pathology may be applied in clinical investigations. Semin. Thromb. Haemost. 2024, 50, 537–551. [Google Scholar] [CrossRef]
  541. Turner, S.; Khan, M.A.; Putrino, D.; Woodcock, A.; Kell, D.B.; Pretorius, E. Long COVID: Pathophysiological factors and abnormal coagulation. Trends Endocrinol. Metab. 2023, 34, 321–344. [Google Scholar] [CrossRef]
  542. Kell, D.B.; Khan, M.A.; Kane, B.; Lip, G.Y.H.; Pretorius, E. Possible role of fibrinaloid microclots in Postural Orthostatic Tachycardia Syndrome (POTS): Focus on Long COVID. J. Pers. Med. 2024, 14, 170. [Google Scholar] [CrossRef] [PubMed]
  543. Fontana, M.; Ćorović, A.; Scully, P.; Moon, J.C. Myocardial Amyloidosis: The Exemplar Interstitial Disease. JACC Cardiovasc. Imaging 2019, 12, 2345–2356. [Google Scholar] [CrossRef]
  544. Pucci, A.; Aimo, A.; Musetti, V.; Barison, A.; Vergaro, G.; Genovesi, D.; Giorgetti, A.; Masotti, S.; Arzilli, C.; Prontera, C.; et al. Amyloid Deposits and Fibrosis on Left Ventricular Endomyocardial Biopsy Correlate With Extracellular Volume in Cardiac Amyloidosis. J. Am. Heart Assoc. 2021, 10, e020358. [Google Scholar] [CrossRef] [PubMed]
  545. Kell, D.B.; Pretorius, E. Proteins behaving badly. Substoichiometric molecular control and amplification of the initiation and nature of amyloid fibril formation: Lessons from and for blood clotting. Progr Biophys. Mol. Biol. 2017, 123, 16–41. [Google Scholar] [CrossRef]
  546. Pretorius, E.; Mbotwe, S.; Bester, J.; Robinson, C.J.; Kell, D.B. Acute induction of anomalous and amyloidogenic blood clotting by molecular amplification of highly substoichiometric levels of bacterial lipopolysaccharide. J. R. Soc. Interface 2016, 123, 20160539. [Google Scholar] [CrossRef]
  547. Pretorius, E.; Page, M.J.; Hendricks, L.; Nkosi, N.B.; Benson, S.R.; Kell, D.B. Both lipopolysaccharide and lipoteichoic acids potently induce anomalous fibrin amyloid formation: Assessment with novel Amytracker™ stains. J. R. Soc. Interface 2018, 15, 20170941. [Google Scholar]
  548. Pretorius, E.; Vlok, M.; Venter, C.; Bezuidenhout, J.A.; Laubscher, G.J.; Steenkamp, J.; Kell, D.B. Persistent clotting protein pathology in Long COVID/Post-Acute Sequelae of COVID-19 (PASC) is accompanied by increased levels of antiplasmin. Cardiovasc. Diabetol. 2021, 20, 172. [Google Scholar]
  549. Appelman, B.; Charlton, B.T.; Goulding, R.P.; Kerkhoff, T.J.; Breedveld, E.A.; Noort, W.; Offringa, C.; Bloemers, F.W.; van Weeghel, M.; Schomakers, B.V.; et al. Muscle abnormalities worsen after post-exertional malaise in long COVID. Nat. Commun. 2024, 15, 17. [Google Scholar] [CrossRef]
  550. Nabai, L.; Ghahary, A.; Jackson, J. Localized Controlled Release of Kynurenic Acid Encapsulated in Synthetic Polymer Reduces Implant-Induced Dermal Fibrosis. Pharmaceutics 2022, 14, 1546. [Google Scholar] [CrossRef]
  551. Papp, A.; Hartwell, R.; Evans, M.; Ghahary, A. The Safety and Tolerability of Topically Delivered Kynurenic Acid in Humans. A Phase 1 Randomized Double-Blind Clinical Trial. J. Pharm. Sci. 2018, 107, 1572–1576. [Google Scholar] [CrossRef]
  552. Poormasjedi-Meibod, M.S.; Pakyari, M.; Jackson, J.K.; Salimi Elizei, S.; Ghahary, A. Development of a nanofibrous wound dressing with an antifibrogenic properties in vitro and in vivo model. J. Biomed. Mater. Res. A 2016, 104, 2334–2344. [Google Scholar] [CrossRef]
  553. Glavin, G.B.; Pinsky, C. Kynurenic acid attenuates experimental ulcer formation and basal gastric acid secretion in rats. Res. Commun. Chem. Pathol. Pharmacol. 1989, 64, 111–119. [Google Scholar]
  554. Glavin, G.B.; Bose, R.; Pinsky, C. Kynurenic acid protects against gastroduodenal ulceration in mice injected with extracts from poisonous Atlantic shellfish. Prog. Neuropsychopharmacol. Biol. Psychiatry 1989, 13, 569–572. [Google Scholar] [CrossRef] [PubMed]
  555. Zhen, D.; Liu, J.; Zhang, X.D.; Song, Z. Kynurenic Acid Acts as a Signaling Molecule Regulating Energy Expenditure and Is Closely Associated With Metabolic Diseases. Front. Endocrinol. 2022, 13, 847611. [Google Scholar] [CrossRef]
  556. Liu, J.J.; Ching, J.; Wee, H.N.; Liu, S.; Gurung, R.L.; Lee, J.; M, Y.; Zheng, H.; Lee, L.S.; Ang, K.; et al. Plasma Tryptophan-Kynurenine Pathway Metabolites and Risk for Progression to End-Stage Kidney Disease in Patients with Type 2 Diabetes. Diabetes Care 2023, 46, 2223–2231. [Google Scholar] [CrossRef] [PubMed]
  557. Bigelman, E.; Pasmanik-Chor, M.; Dassa, B.; Itkin, M.; Malitsky, S.; Dorot, O.; Pichinuk, E.; Kleinberg, Y.; Keren, G.; Entin-Meer, M. Kynurenic acid, a key L-tryptophan-derived metabolite, protects the heart from an ischemic damage. PLoS ONE 2023, 18, e0275550. [Google Scholar] [CrossRef]
  558. Kamel, R.; Baetz, D.; Gueguen, N.; Lebeau, L.; Barbelivien, A.; Guihot, A.L.; Allawa, L.; Gallet, J.; Beaumont, J.; Ovize, M.; et al. Kynurenic Acid: A Novel Player in Cardioprotection against Myocardial Ischemia/Reperfusion Injuries. Pharmaceuticals 2023, 16, 1381. [Google Scholar] [CrossRef] [PubMed]
  559. Bądzyńska, B.; Zakrocka, I.; Sadowski, J.; Turski, W.A.; Kompanowska-Jezierska, E. Effects of systemic administration of kynurenic acid and glycine on renal haemodynamics and excretion in normotensive and spontaneously hypertensive rats. Eur. J. Pharmacol. 2014, 743, 37–41. [Google Scholar] [CrossRef] [PubMed]
  560. Hu, J.; Dai, J.; Sheng, N. Kynurenic Acid Plays a Protective Role in Hepatotoxicity Induced by HFPO-DA in Male Mice. Environ. Sci. Technol. 2024, 58, 1842–1853. [Google Scholar] [CrossRef] [PubMed]
  561. Pyun, D.H.; Kim, T.J.; Kim, M.J.; Hong, S.A.; Abd El-Aty, A.M.; Jeong, J.H.; Jung, T.W. Endogenous metabolite, kynurenic acid, attenuates nonalcoholic fatty liver disease via AMPK/autophagy- and AMPK/ORP150-mediated signaling. J. Cell Physiol. 2021, 236, 4902–4912. [Google Scholar] [CrossRef] [PubMed]
  562. Wang, G.; Cao, K.; Liu, K.; Xue, Y.; Roberts, A.I.; Li, F.; Han, Y.; Rabson, A.B.; Wang, Y.; Shi, Y. Kynurenic acid, an IDO metabolite, controls TSG-6-mediated immunosuppression of human mesenchymal stem cells. Cell Death Differ. 2018, 25, 1209–1223. [Google Scholar] [CrossRef] [PubMed]
  563. Hsieh, Y.C.; Chen, R.F.; Yeh, Y.S.; Lin, M.T.; Hsieh, J.H.; Chen, S.H. Kynurenic acid attenuates multiorgan dysfunction in rats after heatstroke. Acta Pharmacol. Sin. 2011, 32, 167–174. [Google Scholar] [CrossRef] [PubMed]
  564. Balla, Z.; Kormányos, E.S.; Kui, B.; Bálint, E.R.; Für, G.; Orján, E.M.; Iványi, B.; Vécsei, L.; Fülöp, F.; Varga, G.; et al. Kynurenic Acid and Its Analogue SZR-72 Ameliorate the Severity of Experimental Acute Necrotizing Pancreatitis. Front. Immunol. 2021, 12, 702764. [Google Scholar] [CrossRef]
  565. Nahomi, R.B.; Nam, M.H.; Rankenberg, J.; Rakete, S.; Houck, J.A.; Johnson, G.C.; Stankowska, D.L.; Pantcheva, M.B.; MacLean, P.S.; Nagaraj, R.H. Kynurenic Acid Protects Against Ischemia/Reperfusion-Induced Retinal Ganglion Cell Death in Mice. Int. J. Mol. Sci. 2020, 21, 1795. [Google Scholar] [CrossRef] [PubMed]
  566. Poles, M.Z.; Nászai, A.; Gulácsi, L.; Czakó, B.L.; Gál, K.G.; Glenz, R.J.; Dookhun, D.; Rutai, A.; Tallósy, S.P.; Szabo, A.; et al. Kynurenic Acid and Its Synthetic Derivatives Protect Against Sepsis-Associated Neutrophil Activation and Brain Mitochondrial Dysfunction in Rats. Front. Immunol. 2021, 12, 717157. [Google Scholar] [CrossRef]
  567. Moroni, F.; Cozzi, A.; Sili, M.; Mannaioni, G. Kynurenic acid: A metabolite with multiple actions and multiple targets in brain and periphery. J. Neural Transm. 2012, 119, 133–139. [Google Scholar] [CrossRef] [PubMed]
  568. Teunis, C.J.; Stroes, E.S.G.; Boekholdt, S.M.; Wareham, N.J.; Murphy, A.J.; Nieuwdorp, M.; Hazen, S.L.; Hanssen, N.M.J. Tryptophan metabolites and incident cardiovascular disease: The EPIC-Norfolk prospective population study. Atherosclerosis 2023, 387, 117344. [Google Scholar] [CrossRef] [PubMed]
  569. Mangas, A.; Yajeya, J.; González, N.; Ruiz, I.; Pernia, M.; Geffard, M.; Coveñas, R. Gemst: A taylor-made combination that reverts neuroanatomical changes in stroke. Eur. J. Histochem. 2017, 61, 2790. [Google Scholar] [CrossRef]
  570. Mangas, A.; Yajeya, J.; González, N.; Ruiz, I.; Duleu, S.; Geffard, M.; Coveñas, R. Overexpression of kynurenic acid in stroke: An endogenous neuroprotector? Ann. Anat. 2017, 211, 33–38. [Google Scholar] [CrossRef] [PubMed]
  571. Baumgartner, R.; Berg, M.; Matic, L.; Polyzos, K.P.; Forteza, M.J.; Hjorth, S.A.; Schwartz, T.W.; Paulsson-Berne, G.; Hansson, G.K.; Hedin, U.; et al. Evidence that a deviation in the kynurenine pathway aggravates atherosclerotic disease in humans. J. Intern. Med. 2021, 289, 53–68. [Google Scholar] [CrossRef]
  572. Matysik-Woźniak, A.; Turski, W.A.; Turska, M.; Paduch, R.; Łańcut, M.; Piwowarczyk, P.; Czuczwar, M.; Rejdak, R. Kynurenic Acid Accelerates Healing of Corneal Epithelium In Vitro and In Vivo. Pharmaceuticals 2021, 14, 753. [Google Scholar] [CrossRef] [PubMed]
  573. Nestor, M.S.; Berman, B.; Fischer, D.L.; Han, H.; Gade, A.; Arnold, D.; Lawson, A. A Randomized, Double-Blind, Active- and Placebo-Controlled Trial Evaluating a Novel Topical Treatment for Keloid Scars. J. Drugs Dermatol. 2021, 20, 964–968. [Google Scholar] [CrossRef] [PubMed]
  574. Poormasjedi-Meibod, M.S.; Hartwell, R.; Kilani, R.T.; Ghahary, A. Anti-scarring properties of different tryptophan derivatives. PLoS ONE 2014, 9, e91955. [Google Scholar] [CrossRef]
  575. Sorkin, E.M.; Heel, R.C. Guanfacine. A review of its pharmacodynamic and pharmacokinetic properties, and therapeutic efficacy in the treatment of hypertension. Drugs 1986, 31, 301–336. [Google Scholar] [CrossRef]
  576. van Stralen, J.; Gill, S.K.; Reaume, C.J.; Handelman, K. A retrospective medical chart review of clinical outcomes in children and adolescents with attention-deficit/hyperactivity disorder treated with guanfacine extended-release in routine Canadian clinical practice. Child. Adolesc. Psychiatry Ment. Health 2021, 15, 55. [Google Scholar] [CrossRef]
  577. Yu, S.; Shen, S.; Tao, M. Guanfacine for the Treatment of Attention-Deficit Hyperactivity Disorder: An Updated Systematic Review and Meta-Analysis. J. Child. Adolesc. Psychopharmacol. 2023, 33, 40–50. [Google Scholar] [CrossRef]
  578. Arnsten, A.F.T.; Ishizawa, Y.; Xie, Z. Scientific rationale for the use of alpha2A-adrenoceptor agonists in treating neuroinflammatory cognitive disorders. Mol. Psychiatry 2023, 28, 4540–4552. [Google Scholar] [CrossRef]
  579. Kapolka, N.J.; Taghon, G.J.; Rowe, J.B.; Morgan, W.M.; Enten, J.F.; Lambert, N.A.; Isom, D.G. DCyFIR: A high-throughput CRISPR platform for multiplexed G protein-coupled receptor profiling and ligand discovery. Proc. Natl. Acad. Sci. USA 2020, 117, 13117–13126. [Google Scholar] [CrossRef] [PubMed]
  580. DiNatale, B.C.; Murray, I.A.; Schroeder, J.C.; Flaveny, C.A.; Lahoti, T.S.; Laurenzana, E.M.; Omiecinski, C.J.; Perdew, G.H. Kynurenic acid is a potent endogenous aryl hydrocarbon receptor ligand that synergistically induces interleukin-6 in the presence of inflammatory signaling. Toxicol. Sci. 2010, 115, 89–97. [Google Scholar] [CrossRef]
  581. Cuartero, M.I.; de la Parra, J.; Garcia-Culebras, Á.; Ballesteros, I.; Lizasoain, I.; Moro, M.A. The Kynurenine Pathway in the Acute and Chronic Phases of Cerebral Ischemia. Curr. Pharm. Des. 2016, 22, 1060–1073. [Google Scholar] [CrossRef] [PubMed]
  582. Wang, Z.; Yin, L.; Qi, Y.; Zhang, J.; Zhu, H.; Tang, J. Intestinal Flora-Derived Kynurenic Acid Protects Against Intestinal Damage Caused by Candida albicans Infection via Activation of Aryl Hydrocarbon Receptor. Front. Microbiol. 2022, 13, 934786. [Google Scholar] [CrossRef]
  583. Badawy, A.A. Immunotherapy of COVID-19 with poly (ADP-ribose) polymerase inhibitors: Starting with nicotinamide. Biosci. Rep. 2020, 40, BSR20202856. [Google Scholar] [CrossRef]
  584. Siddiqui, T.; Bhattarai, P.; Popova, S.; Cosacak, M.I.; Sariya, S.; Zhang, Y.; Mayeux, R.; Tosto, G.; Kizil, C. KYNA/Ahr Signaling Suppresses Neural Stem Cell Plasticity and Neurogenesis in Adult Zebrafish Model of Alzheimer’s Disease. Cells 2021, 10, 2748. [Google Scholar] [CrossRef]
  585. Shi, Y.; Zeng, Z.; Yu, J.; Tang, B.; Tang, R.; Xiao, R. The aryl hydrocarbon receptor: An environmental effector in the pathogenesis of fibrosis. Pharmacol. Res. 2020, 160, 105180. [Google Scholar] [CrossRef] [PubMed]
  586. Fernández-Gallego, N.; Sánchez-Madrid, F.; Cibrian, D. Role of AHR Ligands in Skin Homeostasis and Cutaneous Inflammation. Cells 2021, 10, 3176. [Google Scholar] [CrossRef]
  587. Dadvar, S.; Ferreira, D.M.S.; Cervenka, I.; Ruas, J.L. The weight of nutrients: Kynurenine metabolites in obesity and exercise. J. Intern. Med. 2018, 284, 519–533. [Google Scholar] [CrossRef]
  588. Wyant, G.A.; Yu, W.; Doulamis, I.P.; Nomoto, R.S.; Saeed, M.Y.; Duignan, T.; McCully, J.D.; Kaelin, W.G., Jr. Mitochondrial remodeling and ischemic protection by G protein-coupled receptor 35 agonists. Science 2022, 377, 621–629. [Google Scholar] [CrossRef]
  589. Milligan, G. Orthologue selectivity and ligand bias: Translating the pharmacology of GPR35. Trends Pharmacol. Sci. 2011, 32, 317–325. [Google Scholar] [CrossRef]
  590. Mackenzie, A.E.; Lappin, J.E.; Taylor, D.L.; Nicklin, S.A.; Milligan, G. GPR35 as a Novel Therapeutic Target. Front. Endocrinol. 2011, 2, 68. [Google Scholar] [CrossRef]
  591. Wu, Y.; Zhang, P.; Fan, H.; Zhang, C.; Yu, P.; Liang, X.; Chen, Y. GPR35 acts a dual role and therapeutic target in inflammation. Front. Immunol. 2023, 14, 1254446. [Google Scholar] [CrossRef]
  592. Ahmadi, S.; Muth-Selbach, U.; Lauterbach, A.; Lipfert, P.; Neuhuber, W.L.; Zeilhofer, H.U. Facilitation of spinal NMDA receptor currents by spillover of synaptically released glycine. Science 2003, 300, 2094–2097. [Google Scholar] [CrossRef] [PubMed]
  593. Parsons, C.G.; Danysz, W.; Quack, G.; Hartmann, S.; Lorenz, B.; Wollenburg, C.; Baran, L.; Przegalinski, E.; Kostowski, W.; Krzascik, P.; et al. Novel systemically active antagonists of the glycine site of the N-methyl-D-aspartate receptor: Electrophysiological, biochemical and behavioral characterization. J. Pharmacol. Exp. Ther. 1997, 283, 1264–1275. [Google Scholar] [PubMed]
  594. Schwarcz, R.; Stone, T.W. The kynurenine pathway and the brain: Challenges, controversies and promises. Neuropharmacology 2017, 112, 237–247. [Google Scholar] [CrossRef] [PubMed]
  595. Lemieux, G.A.; Cunningham, K.A.; Lin, L.; Mayer, F.; Werb, Z.; Ashrafi, K. Kynurenic acid is a nutritional cue that enables behavioral plasticity. Cell 2015, 160, 119–131. [Google Scholar] [CrossRef]
  596. Bagasrawala, I.; Zecevic, N.; Radonjić, N.V. N-Methyl D-Aspartate Receptor Antagonist Kynurenic Acid Affects Human Cortical Development. Front. Neurosci. 2016, 10, 435. [Google Scholar] [CrossRef]
  597. Mok, M.H.S.; Fricker, A.C.; Weil, A.; Kew, J.N.C. Electrophysiological characterisation of the actions of kynurenic acid at ligand-gated ion channels. Neuropharmacology 2009, 57, 242–249. [Google Scholar] [CrossRef]
  598. Erhardt, S.; Olsson, S.K.; Engberg, G. Pharmacological manipulation of kynurenic acid: Potential in the treatment of psychiatric disorders. CNS Drugs 2009, 23, 91–101. [Google Scholar] [CrossRef]
  599. Stone, T.W. Does kynurenic acid act on nicotinic receptors? An assessment of the evidence. J. Neurochem. 2020, 152, 627–649. [Google Scholar] [CrossRef]
  600. Chen, Y.; Guillemin, G.J. Kynurenine pathway metabolites in humans: Disease and healthy States. Int. J. Tryptophan Res. 2009, 2, 1–19. [Google Scholar] [CrossRef] [PubMed]
  601. Hilmas, C.; Pereira, E.F.R.; Alkondon, M.; Rassoulpour, A.; Schwarcz, R.; Albuquerque, E.X. The brain metabolite kynurenic acid inhibits alpha7 nicotinic receptor activity and increases non-alpha7 nicotinic receptor expression: Physiopathological implications. J. Neurosci. 2001, 21, 7463–7473. [Google Scholar] [CrossRef] [PubMed]
  602. Dobelis, P.; Staley, K.J.; Cooper, D.C. Lack of modulation of nicotinic acetylcholine alpha-7 receptor currents by kynurenic acid in adult hippocampal interneurons. PLoS ONE 2012, 7, e41108. [Google Scholar] [CrossRef]
  603. Steiner, L.; Gold, M.; Mengel, D.; Dodel, R.; Bach, J.P. The endogenous alpha7 nicotinic acetylcholine receptor antagonist kynurenic acid modulates amyloid-beta-induced inflammation in BV-2 microglial cells. J. Neurol. Sci. 2014, 344, 94–99. [Google Scholar] [CrossRef] [PubMed]
  604. Torosyan, R.; Huang, S.; Bommi, P.V.; Tiwari, R.; An, S.Y.; Schonfeld, M.; Rajendran, G.; Kavanaugh, M.A.; Gibbs, B.; Truax, A.D.; et al. Hypoxic preconditioning protects against ischemic kidney injury through the IDO1/kynurenine pathway. Cell Rep. 2021, 36, 109547. [Google Scholar] [CrossRef] [PubMed]
  605. Wu, M.Y.; Yiang, G.T.; Liao, W.T.; Tsai, A.P.; Cheng, Y.L.; Cheng, P.W.; Li, C.Y.; Li, C.J. Current Mechanistic Concepts in Ischemia and Reperfusion Injury. Cell Physiol. Biochem. 2018, 46, 1650–1667. [Google Scholar] [CrossRef] [PubMed]
  606. Enzmann, G.; Kargaran, S.; Engelhardt, B. Ischemia-reperfusion injury in stroke: Impact of the brain barriers and brain immune privilege on neutrophil function. Ther. Adv. Neurol. Disord. 2018, 11, 1756286418794184. [Google Scholar] [CrossRef] [PubMed]
  607. Wu, M.; Gu, X.; Ma, Z. Mitochondrial Quality Control in Cerebral Ischemia-Reperfusion Injury. Mol. Neurobiol. 2021, 58, 5253–5271. [Google Scholar] [CrossRef] [PubMed]
  608. Hausenloy, D.J.; Yellon, D.M. Myocardial ischemia-reperfusion injury: A neglected therapeutic target. J. Clin. Invest. 2013, 123, 92–100. [Google Scholar] [CrossRef] [PubMed]
  609. Daiber, A.; Andreadou, I.; Oelze, M.; Davidson, S.M.; Hausenloy, D.J. Discovery of new therapeutic redox targets for cardioprotection against ischemia/reperfusion injury and heart failure. Free Radic. Biol. Med. 2021, 163, 325–343. [Google Scholar] [CrossRef] [PubMed]
  610. Murphy, E.; Steenbergen, C. Mechanisms underlying acute protection from cardiac ischemia-reperfusion injury. Physiol. Rev. 2008, 88, 581–609. [Google Scholar] [CrossRef]
  611. Aigner, F.; Maier, H.T.; Schwelberger, H.G.; Wallnofer, E.A.; Amberger, A.; Obrist, P.; Berger, T.; Mak, T.W.; Maglione, M.; Margreiter, R.; et al. Lipocalin-2 regulates the inflammatory response during ischemia and reperfusion of the transplanted heart. Amer J. Transplant. 2007, 7, 779–788. [Google Scholar]
  612. Jaeschke, H.; Woolbright, B.L. Current strategies to minimize hepatic ischemia-reperfusion injury by targeting reactive oxygen species. Transplant. Rev. 2012, 26, 103–114. [Google Scholar] [CrossRef]
  613. Situmorang, G.R.; Sheerin, N.S. Ischaemia reperfusion injury: Mechanisms of progression to chronic graft dysfunction. Pediatr. Nephrol. 2019, 34, 951–963. [Google Scholar] [CrossRef]
  614. Zhai, Y.; Petrowsky, H.; Hong, J.C.; Busuttil, R.W.; Kupiec-Weglinski, J.W. Ischaemia-reperfusion injury in liver transplantation--from bench to bedside. Nat. Rev. Gastroenterol. Hepatol. 2013, 10, 79–89. [Google Scholar] [CrossRef]
  615. Williamson, R.D.; McCarthy, F.P.; Manna, S.; Groarke, E.; Kell, D.B.; Kenny, L.C.; McCarthy, C.M. L-(+)-Ergothioneine Significantly Improves the Clinical Characteristics of Preeclampsia in the Reduced Uterine Perfusion Pressure Rat Model. Hypertension 2020, 75, 561–568. [Google Scholar] [CrossRef] [PubMed]
  616. Taracha, E.; Czarna, M.; Turzynska, D.; Sobolewska, A.; Maciejak, P. Long-term disruption of tissue levels of glutamate and glutamatergic neurotransmission neuromodulators, taurine and kynurenic acid induced by amphetamine. Psychopharmacology 2024, 241, 1387–1398. [Google Scholar] [CrossRef] [PubMed]
  617. Marques, C.; Fernandes, I.; Meireles, M.; Faria, A.; Spencer, J.P.E.; Mateus, N.; Calhau, C. Gut microbiota modulation accounts for the neuroprotective properties of anthocyanins. Sci. Rep. 2018, 8, 11341. [Google Scholar] [CrossRef]
  618. Bednarz, K.; Kozieł, K.; Urbańska, E.M. Novel Activity of Oral Hypoglycemic Agents Linked with Decreased Formation of Tryptophan Metabolite, Kynurenic Acid. Life 2024, 14, 127. [Google Scholar] [CrossRef]
  619. Liang, Y.Y.; Zhang, L.D.; Luo, X.; Wu, L.L.; Chen, Z.W.; Wei, G.H.; Zhang, K.Q.; Du, Z.A.; Li, R.Z.; So, K.F.; et al. All roads lead to Rome—A review of the potential mechanisms by which exerkines exhibit neuroprotective effects in Alzheimer’s disease. Neural Regen. Res. 2022, 17, 1210–1227. [Google Scholar] [CrossRef] [PubMed]
  620. Bian, X.; Wang, Q.; Wang, Y.; Lou, S. The function of previously unappreciated exerkines secreted by muscle in regulation of neurodegenerative diseases. Front. Mol. Neurosci. 2023, 16, 1305208. [Google Scholar] [CrossRef]
  621. Lim, A.; Harijanto, C.; Vogrin, S.; Guillemin, G.; Duque, G. Does Exercise Influence Kynurenine/Tryptophan Metabolism and Psychological Outcomes in Persons With Age-Related Diseases? A Systematic Review. Int. J. Tryptophan Res. 2021, 14, 1178646921991119. [Google Scholar] [CrossRef]
  622. Joisten, N.; Walzik, D.; Schenk, A.; Metcalfe, A.J.; Belen, S.; Schaaf, K.; Jacko, D.; Gehlert, S.; Spiliopoulou, P.; Garzinsky, A.M.; et al. Acute exercise activates the AHR in peripheral blood mononuclear cells in an intensity-dependent manner. Am. J. Physiol. Cell Physiol. 2024, 327, C438–C445. [Google Scholar] [CrossRef] [PubMed]
  623. Juhas, U.; Reczkowicz, J.; Kortas, J.A.; Zychowska, M.; Pilis, K.; Ziemann, E.; Cytrych, I.; Antosiewicz, J.; Borkowska, A. Eight-day fasting modulates serum kynurenines in healthy men at rest and after exercise. Front. Endocrinol. 2024, 15, 1403491. [Google Scholar] [CrossRef]
  624. Louvrou, V.; Solianik, R.; Brazaitis, M.; Erhardt, S. Exploring the effect of prolonged fasting on kynurenine pathway metabolites and stress markers in healthy male individuals. Eur. J. Clin. Nutr. 2024. [Google Scholar] [CrossRef]
  625. Tomczyk, T.; Urbańska, E.M. Experimental hypothyroidism raises brain kynurenic acid—Novel aspect of thyroid dysfunction. Eur. J. Pharmacol. 2020, 883, 173363. [Google Scholar] [CrossRef]
  626. Guillemin, G.J.; Smith, D.G.; Kerr, S.J.; Smythe, G.A.; Kapoor, V.; Armati, P.J.; Brew, B.J. Characterisation of kynurenine pathway metabolism in human astrocytes and implications in neuropathogenesis. Redox Rep. 2000, 5, 108–111. [Google Scholar] [CrossRef]
  627. Guillemin, G.J.; Cullen, K.M.; Lim, C.K.; Smythe, G.A.; Garner, B.; Kapoor, V.; Takikawa, O.; Brew, B.J. Characterization of the kynurenine pathway in human neurons. J. Neurosci. 2007, 27, 12884–12892. [Google Scholar] [CrossRef]
  628. Żarnowski, T.; Chorągiewicz, T.; Tulidowicz-Bielak, M.; Thaler, S.; Rejdak, R.; Żarnowski, I.; Turski, W.A.; Gasior, M. Ketogenic diet increases concentrations of kynurenic acid in discrete brain structures of young and adult rats. J. Neural Transm. 2012, 119, 679–684. [Google Scholar] [CrossRef]
  629. Żarnowski, T.; Tulidowicz-Bielak, M.; Żarnowska, I.; Mitosek-Szewczyk, K.; Wnorowski, A.; Jozwiak, K.; Gasior, M.; Turski, W.A. Kynurenic Acid and Neuroprotective Activity of the Ketogenic Diet in the Eye. Curr. Med. Chem. 2017, 24, 3547–3558. [Google Scholar] [CrossRef] [PubMed]
  630. Kotańska, M.; Łanocha, M.; Bednarski, M.; Marcinkowska, M. MM165—A Small Hybrid Molecule Modulates the Kynurenine Pathway and Attenuates Lipopolysaccharide-Induced Memory Deficits and Inflammation. Neurochem. Res. 2024, 49, 1200–1211. [Google Scholar] [CrossRef]
  631. Kupjetz, M.; Patt, N.; Joisten, N.; Ueland, P.M.; McCann, A.; Gonzenbach, R.; Bansi, J.; Zimmer, P. The serum kynurenine pathway metabolic profile is associated with overweight and obesity in multiple sclerosis. Mult. Scler. Relat. Disord. 2023, 72, 104592. [Google Scholar] [CrossRef]
  632. Bjørke-Monsen, A.L.; Varsi, K.; Sakkestad, S.T.; Ulvik, A.; Ebbing, C.; Ueland, P.M. Lower levels of the neuroprotective tryptophan metabolite, kynurenic acid, in users of estrogen contraceptives. Sci. Rep. 2023, 13, 16370. [Google Scholar] [CrossRef]
  633. de Bie, J.; Lim, C.K.; Guillemin, G.J. Progesterone Alters Kynurenine Pathway Activation in IFN-gamma-Activated Macrophages—Relevance for Neuroinflammatory Diseases. Int. J. Tryptophan Res. 2016, 9, 89–93. [Google Scholar] [CrossRef] [PubMed]
  634. Wu, L.; Hu, Z.; Lv, Y.; Ge, C.; Luo, X.; Zhan, S.; Huang, W.; Shen, X.; Yu, D.; Liu, B. Hericium erinaceus polysaccharides ameliorate nonalcoholic fatty liver disease via gut microbiota and tryptophan metabolism regulation in an aged laying hen model. Int. J. Biol. Macromol. 2024, 273, 132735. [Google Scholar] [CrossRef]
  635. Klausing, A.D.; Fukuwatari, T.; Bucci, D.J.; Schwarcz, R. Stress-induced impairment in fear discrimination is causally related to increased kynurenic acid formation in the prefrontal cortex. Psychopharmacology 2020, 237, 1931–1941. [Google Scholar] [CrossRef]
  636. Yan, L.; Wang, W.J.; Cheng, T.; Yang, D.R.; Wang, Y.J.; Wang, Y.Z.; Yang, F.Z.; So, K.F.; Zhang, L. Hepatic kynurenic acid mediates phosphorylation of Nogo-A in the medial prefrontal cortex to regulate chronic stress-induced anxiety-like behaviors in mice. Acta Pharmacol. Sin. 2024. Online ahead of print. [Google Scholar] [CrossRef]
  637. Deora, G.S.; Kantham, S.; Chan, S.; Dighe, S.N.; Veliyath, S.K.; McColl, G.; Parat, M.O.; McGeary, R.P.; Ross, B.P. Multifunctional Analogs of Kynurenic Acid for the Treatment of Alzheimer’s Disease: Synthesis, Pharmacology, and Molecular Modeling Studies. ACS Chem. Neurosci. 2017, 8, 2667–2675. [Google Scholar] [CrossRef]
  638. Szabo, M.; Lajkó, N.; Dulka, K.; Szatmári, I.; Fülöp, F.; Mihály, A.; Vécsei, L.; Gulya, K. Kynurenic Acid and Its Analog SZR104 Exhibit Strong Antiinflammatory Effects and Alter the Intracellular Distribution and Methylation Patterns of H3 Histones in Immunochallenged Microglia-Enriched Cultures of Newborn Rat Brains. Int. J. Mol. Sci. 2022, 23, 1079. [Google Scholar] [CrossRef]
  639. Baris, E.; Simsek, O.; Yoca, O.U.; Demir, A.B.; Tosun, M. Effects of kynurenic acid and choline on lipopolysaccharide-induced cyclooxygenase pathway. Turk. J. Biochem. 2023, 48, 311–318. [Google Scholar] [CrossRef]
  640. Małaczewska, J.; Siwicki, A.K.; Wócik, R.M.; Kaczorek, E.; Turski, W.A. Effect of oral administration of kynurenic acid on the activity of the peripheral blood leukocytes in mice. Cent. Eur. J. Immunol. 2014, 39, 6–13. [Google Scholar] [CrossRef] [PubMed]
  641. Małaczewska, J.; Siwicki, A.K.; Wójcik, R.M.; Turski, W.A.; Kaczorek, E. The effect of kynurenic acid on the synthesis of selected cytokines by murine splenocytes—in vitro and ex vivo studies. Cent. Eur. J. Immunol. 2016, 41, 39–46. [Google Scholar] [CrossRef] [PubMed]
  642. Lajkó, N.; Kata, D.; Szabó, M.; Mátyás, A.; Dulka, K.; Földesi, I.; Fülöp, F.; Gulya, K.; Vécsei, L.; Mihály, A. Sensitivity of Rodent Microglia to Kynurenines in Models of Epilepsy and Inflammation In Vivo and In Vitro: Microglia Activation is Inhibited by Kynurenic Acid and the Synthetic Analogue SZR104. Int. J. Mol. Sci. 2020, 21, 9333. [Google Scholar] [CrossRef] [PubMed]
  643. Csáti, A.; Edvinsson, L.; Vécsei, L.; Toldi, J.; Fülöp, F.; Tajti, J.; Warfvinge, K. Kynurenic acid modulates experimentally induced inflammation in the trigeminal ganglion. J. Headache Pain 2015, 16, 99. [Google Scholar] [CrossRef] [PubMed]
  644. Zhang, Y.W.; Pang, X.; Yang, Y. Hydrogels containing KYNA promote angiogenesis and inhibit inflammation to improve the survival rate of multi-territory perforator flaps. Biomed. Pharmacother. 2024, 174, 116454. [Google Scholar] [CrossRef]
  645. Robinson, K.S.L.; Stewart, A.M.; Cachat, J.; Landsman, S.; Gebhardt, M.; Kalueff, A.V. Psychopharmacological effects of acute exposure to kynurenic acid (KYNA) in zebrafish. Pharmacol. Biochem. Behav. 2013, 108, 54–60. [Google Scholar] [CrossRef] [PubMed]
  646. Kim, H.H.; Jeong, S.H.; Ha, S.E.; Park, M.Y.; Bhosale, P.B.; Abusaliya, A.; Won, C.K.; Heo, J.D.; Kim, H.W.; Kim, G.S. Cellular Regulation of Kynurenic Acid-Induced Cell Apoptosis Pathways in AGS Cells. Int. J. Mol. Sci. 2022, 23, 8894. [Google Scholar] [CrossRef] [PubMed]
  647. Kim, H.H.; Ha, S.E.; Park, M.Y.; Jeong, S.H.; Bhosale, P.B.; Abusaliya, A.; Won, C.K.; Heo, J.D.; Ahn, M.; Seong, J.K.; et al. Identification of Kynurenic Acid-Induced Apoptotic Biomarkers in Gastric Cancer-Derived AGS Cells through Next-Generation Transcriptome Sequencing Analysis. Nutrients 2023, 15, 193. [Google Scholar] [CrossRef]
  648. Lun, J.; Li, Y.; Gao, X.; Gong, Z.; Chen, X.; Zou, J.; Zhou, C.; Huang, Y.; Zhou, B.; Huang, P.; et al. Kynurenic acid blunts A1 astrocyte activation against neurodegeneration in HIV-associated neurocognitive disorders. J. Neuroinflamm. 2023, 20, 87. [Google Scholar] [CrossRef]
  649. Młotkowska, P.; Misztal, T.; Kowalczyk, P.; Marciniak, E. Effect of kynurenic acid on enzymatic activity of the DNA base excision repair pathway in specific areas of the sheep brain. Sci. Rep. 2024, 14, 15506. [Google Scholar] [CrossRef]
  650. Wejksza, K.; Rzeski, W.; Turski, W.A. Kynurenic acid protects against the homocysteine-induced impairment of endothelial cells. Pharmacol. Rep. 2009, 61, 751–756. [Google Scholar] [CrossRef]
  651. Lajkó, E.; Tuka, B.; Fülöp, F.; Krizbai, I.; Toldi, J.; Magyar, K.; Vécsei, L.; Köhidai, L. Kynurenic acid and its derivatives are able to modulate the adhesion and locomotion of brain endothelial cells. J. Neural Transm. 2018, 125, 899–912. [Google Scholar] [CrossRef]
  652. Di Serio, C.; Cozzi, A.; Angeli, I.; Doria, L.; Micucci, I.; Pellerito, S.; Mirone, P.; Masotti, G.; Moroni, F.; Tarantini, F. Kynurenic acid inhibits the release of the neurotrophic fibroblast growth factor (FGF)-1 and enhances proliferation of glia cells, in vitro. Cell Mol. Neurobiol. 2005, 25, 981–993. [Google Scholar] [CrossRef] [PubMed]
  653. Wu, H.Q.; Pereira, E.F.; Bruno, J.P.; Pellicciari, R.; Albuquerque, E.X.; Schwarcz, R. The astrocyte-derived alpha7 nicotinic receptor antagonist kynurenic acid controls extracellular glutamate levels in the prefrontal cortex. J. Mol. Neurosci. 2010, 40, 204–210. [Google Scholar] [CrossRef]
  654. Bądzyńska, B.; Zakrocka, I.; Turski, W.A.; Olszyński, K.H.; Sadowski, J.; Kompanowska-Jezierska, E. Kynurenic acid selectively reduces heart rate in spontaneously hypertensive rats. Naunyn Schmiedebergs Arch. Pharmacol. 2020, 393, 673–679. [Google Scholar] [CrossRef] [PubMed]
  655. Ye, Y.; Zhang, X.; Su, D.; Ren, Y.; Cheng, F.; Yao, Y.; Shi, G.; Ji, Y.; Chen, S.; Shi, P.; et al. Therapeutic efcacy of human adipose mesenchymal stem cells in Crohn’s colon fbrosis is improved by IFN-γ and kynurenic acid priming through indoleamine 2,3-dioxygenase-1 signaling. Cell Res. Ther. 2022, 13, 465. [Google Scholar] [CrossRef]
  656. Jung, T.W.; Park, J.; Sun, J.L.; Ahn, S.H.; Abd El-Aty, A.M.; Hacimuftuoglu, A.; Kim, H.C.; Shim, J.H.; Shin, S.; Jeong, J.H. Administration of kynurenic acid reduces hyperlipidemia-induced inflammation and insulin resistance in skeletal muscle and adipocytes. Mol. Cell Endocrinol. 2020, 518, 110928. [Google Scholar] [CrossRef] [PubMed]
  657. Salimi Elizei, S.; Poormasjedi-Meibod, M.S.; Wang, X.; Kheirandish, M.; Ghahary, A. Kynurenic acid downregulates IL-17/1L-23 axis in vitro. Mol. Cell Biochem. 2017, 431, 55–65. [Google Scholar] [CrossRef]
  658. Klein, C.; Patte-Mensah, C.; Taleb, O.; Bourguignon, J.J.; Schmitt, M.; Bihel, F.; Maitre, M.; Mensah-Nyagan, A.G. The neuroprotector kynurenic acid increases neuronal cell survival through neprilysin induction. Neuropharmacology 2013, 70, 254–260. [Google Scholar] [CrossRef]
  659. Tiszlavicz, Z.; Németh, B.; Fülöp, F.; Vécsei, L.; Tápai, K.; Ocsovszky, I.; Mándi, Y. Different inhibitory effects of kynurenic acid and a novel kynurenic acid analogue on tumour necrosis factor-alpha (TNF-alpha) production by mononuclear cells, HMGB1 production by monocytes and HNP1-3 secretion by neutrophils. Naunyn Schmiedebergs Arch. Pharmacol. 2011, 383, 447–455. [Google Scholar] [CrossRef]
  660. Mándi, Y.; Endrész, V.; Mosolygo, T.; Burián, K.; Lantos, I.; Fülöp, F.; Szatmári, I.; Lörinczi, B.; Balog, A.; Vécsei, L. The Opposite Effects of Kynurenic Acid and Different Kynurenic Acid Analogs on Tumor Necrosis Factor-alpha (TNF-alpha) Production and Tumor Necrosis Factor-Stimulated Gene-6 (TSG-6) Expression. Front. Immunol. 2019, 10, 1406. [Google Scholar] [CrossRef]
  661. Balog, A.; Varga, B.; Fülöp, F.; Lantos, I.; Toldi, G.; Vécsei, L.; Mándi, Y. Kynurenic Acid Analog Attenuates the Production of Tumor Necrosis Factor-alpha, Calgranulins (S100A 8/9 and S100A 12), and the Secretion of HNP1-3 and Stimulates the Production of Tumor Necrosis Factor-Stimulated Gene-6 in Whole Blood Cultures of Patients With Rheumatoid Arthritis. Front. Immunol. 2021, 12, 632513. [Google Scholar] [CrossRef]
  662. Joshi, P.; Perni, M.; Limbocker, R.; Mannini, B.; Casford, S.; Chia, S.; Habchi, J.; Labbadia, J.; Dobson, C.M.; Vendruscolo, M. Two human metabolites rescue a C. elegans model of Alzheimer’s disease via a cytosolic unfolded protein response. Commun. Biol. 2021, 4, 843. [Google Scholar] [CrossRef]
  663. Wang, H.; Li, J.; Wang, Z.; Tian, Y.; Li, C.; Jin, F.; Li, J.; Wang, L. Perivascular brown adipocytes-derived kynurenic acid relaxes blood vessel via endothelium PI3K-Akt-eNOS pathway. Biomed. Pharmacother. 2022, 150, 113040. [Google Scholar] [CrossRef]
  664. Hare, S.M.; Adhikari, B.M.; Mo, C.; Chen, S.; Wijtenburg, S.A.; Seneviratne, C.; Kane-Gerard, S.; Sathyasaikumar, K.V.; Notarangelo, F.M.; Schwarcz, R.; et al. Tryptophan challenge in individuals with schizophrenia and healthy controls: Acute effects on circulating kynurenine and kynurenic acid, cognition and cerebral blood flow. Neuropsychopharmacology 2023, 48, 1594–1601. [Google Scholar] [CrossRef]
  665. Turski, W.A.; Malaczewska, J.; Marciniak, S.; Bednarski, J.; Turski, M.P.; Jablonski, M.; Siwicki, A.K. On the toxicity of kynurenic acid in vivo and in vitro. Pharmacol. Rep. 2014, 66, 1127–1133. [Google Scholar] [CrossRef] [PubMed]
  666. Marciniak, S.; Wnorowski, A.; Smolinska, K.; Walczyna, B.; Turski, W.; Kocki, T.; Paluszkiewicz, P.; Parada-Turska, J. Kynurenic Acid Protects against Thioacetamide-Induced Liver Injury in Rats. Anal. Cell Pathol. 2018, 2018, 1270483. [Google Scholar] [CrossRef]
  667. Loikas, P.; Hilakivi, I. Effects of kynurenic acid and ketamine on neonatal sleep in rats. Pharmacol. Toxicol. 1989, 64, 185–189. [Google Scholar] [CrossRef]
  668. Bespalov, A.; Dumpis, M.; Piotrovsky, L.; Zvartau, E. Excitatory amino acid receptor antagonist kynurenic acid attenuates rewarding potential of morphine. Eur. J. Pharmacol. 1994, 264, 233–239. [Google Scholar] [CrossRef]
  669. Majerova, P.; Olesova, D.; Golisova, G.; Buralova, M.; Michalicova, A.; Vegh, J.; Piestansky, J.; Bhide, M.; Hanes, J.; Kovac, A. Analog of kynurenic acid decreases tau pathology by modulating astrogliosis in rat model for tauopathy. Biomed. Pharmacother. 2022, 152, 113257. [Google Scholar] [CrossRef]
  670. Okuno, A.; Fukuwatari, T.; Shibata, K. Urinary excretory ratio of anthranilic acid/kynurenic acid as an index of the tolerable amount of tryptophan. Biosci. Biotechnol. Biochem. 2008, 72, 1667–1672. [Google Scholar] [CrossRef]
  671. Okuno, A.; Fukuwatari, T.; Shibata, K. High tryptophan diet reduces extracellular dopamine release via kynurenic acid production in rat striatum. J. Neurochem. 2011, 118, 796–805. [Google Scholar] [CrossRef] [PubMed]
  672. Hiratsuka, C.; Fukuwatari, T.; Sano, M.; Saito, K.; Sasaki, S.; Shibata, K. Supplementing healthy women with up to 5.0 g/d of L-tryptophan has no adverse effects. J. Nutr. 2013, 143, 859–866. [Google Scholar] [CrossRef] [PubMed]
  673. Hiratsuka, C.; Sano, M.; Fukuwatari, T.; Shibata, K. Time-dependent effects of L-tryptophan administration on urinary excretion of L-tryptophan metabolites. J. Nutr. Sci. Vitaminol. 2014, 60, 255–260. [Google Scholar] [CrossRef]
  674. Al-Karagholi, M.A.; Hansen, J.M.; Abou-Kassem, D.; Hansted, A.K.; Ubhayasekera, K.; Bergquist, J.; Vecsei, L.; Jansen-Olesen, I.; Ashina, M. Phase 1 study to access safety, tolerability, pharmacokinetics, and pharmacodynamics of kynurenine in healthy volunteers. Pharmacol. Res. Perspect. 2021, 9, e00741. [Google Scholar] [CrossRef]
  675. Jauch, D.A.; Sethy, V.H.; Weick, B.G.; Chase, T.N.; Schwarcz, R. Intravenous administration of L-kynurenine to rhesus monkeys: Effect on quinolinate and kynurenate levels in serum and cerebrospinal fluid. Neuropharmacology 1993, 32, 467–472. [Google Scholar] [CrossRef] [PubMed]
  676. Mestres, J.; Gregori-Puigjané, E.; Valverde, S.; Solé, R.V. The topology of drug-target interaction networks: Implicit dependence on drug properties and target families. Mol. Biosyst. 2009, 5, 1051–1057. [Google Scholar]
  677. Hokari, M.; Wu, H.Q.; Schwarcz, R.; Smith, Q.R. Facilitated brain uptake of 4-chlorokynurenine and conversion to 7-chlorokynurenic acid. Neuroreport 1996, 8, 15–18. [Google Scholar] [CrossRef]
  678. Wallace, M.; White, A.; Grako, K.A.; Lane, R.; Cato, A.J.; Snodgrass, H.R. Randomized, double-blind, placebo-controlled, dose-escalation study: Investigation of the safety, pharmacokinetics, and antihyperalgesic activity of l-4-chlorokynurenine in healthy volunteers. Scand. J. Pain. 2017, 17, 243–251. [Google Scholar] [CrossRef]
  679. Park, L.T.; Kadriu, B.; Gould, T.D.; Zanos, P.; Greenstein, D.; Evans, J.W.; Yuan, P.; Farmer, C.A.; Oppenheimer, M.; George, J.M.; et al. A Randomized Trial of the N-Methyl-d-Aspartate Receptor Glycine Site Antagonist Prodrug 4-Chlorokynurenine in Treatment-Resistant Depression. Int. J. Neuropsychopharmacol. 2020, 23, 417–425. [Google Scholar] [CrossRef] [PubMed]
  680. Williams, R.J. Biochemical Individuality; John Wiley: New York, NY, USA, 1956. [Google Scholar]
  681. Ioannidis, J.P.A.; Trikalinos, T.A. Early extreme contradictory estimates may appear in published research: The Proteus phenomenon in molecular genetics research and randomized trials. J. Clin. Epidemiol. 2005, 58, 543–549. [Google Scholar]
  682. Ioannidis, J.P.A. Why most published research findings are false. PLoS Med. 2005, 2, e124. [Google Scholar]
  683. Ioannidis, J.P.A. Contradicted and initially stronger effects in highly cited clinical research. JAMA 2005, 294, 218–228. [Google Scholar]
  684. Büki, A.; Kekesi, G.; Horvath, G.; Vécsei, L. A Potential Interface between the Kynurenine Pathway and Autonomic Imbalance in Schizophrenia. Int. J. Mol. Sci. 2021, 22, 10016. [Google Scholar] [CrossRef] [PubMed]
  685. Santini, A.; Cammarata, S.M.; Capone, G.; Ianaro, A.; Tenore, G.C.; Pani, L.; Novellino, E. Nutraceuticals: Opening the debate for a regulatory framework. Br. J. Clin. Pharmacol. 2018, 84, 659–672. [Google Scholar] [CrossRef]
  686. Wiggins, A.K.A.; Grantham, A.; Anderson, G.H. Optimizing foods for special dietary use in Canada: Key outcomes and recommendations from a tripartite workshop. Appl. Physiol. Nutr. Metab. 2019, 44, 1258–1265. [Google Scholar] [CrossRef]
  687. Bansal, R.; Dhiman, A. Nutraceuticals: A Comparative Analysis of Regulatory Framework in Different Countries of the World. Endocr. Metab. Immune Disord. Drug Targets 2020, 20, 1654–1663. [Google Scholar] [CrossRef]
  688. Blaze, J. A Comparison of Current Regulatory Frameworks for Nutraceuticals in Australia, Canada, Japan, and the United States. Innov. Pharm. 2021, 12. [Google Scholar] [CrossRef]
  689. Thakur, S.; Gupta, M.M.; Sharma, D. Nutraceuticals regulation: An overview of the regulatory frameworks in USA, EU, and Japan. In Nutraceutical Fruits and Foods for Neurodegenerative Disorders; Elsevier: Amsterdam, The Netherlands, 2023; pp. 421–440. [Google Scholar]
  690. Oliver, S.G.; Winson, M.K.; Kell, D.B.; Baganz, F. Systematic functional analysis of the yeast genome. Trends Biotechnol. 1998, 16, 373–378. [Google Scholar] [CrossRef] [PubMed]
  691. Kell, D.B.; Oliver, S.G. The metabolome 18 years on: A concept comes of age. Metabolomics 2016, 12, 148. [Google Scholar] [CrossRef] [PubMed]
  692. Dunn, W.B.; Broadhurst, D.; Begley, P.; Zelena, E.; Francis-McIntyre, S.; Anderson, N.; Brown, N.; Knowles, J.; Halsall, A.; Haselden, J.N.; et al. Procedures for large-scale metabolic profiling of serum and plasma using gas chromatography and liquid chromatography coupled to mass spectrometry. Nat. Protoc. 2011, 6, 1060–1083. [Google Scholar] [CrossRef]
  693. Dunn, W.B.; Erban, A.; Weber, R.J.M.; Creek, D.J.; Brown, M.; Breitling, R.; Hankemeier, T.; Goodacre, R.; Neumann, S.; Kopka, J.; et al. Mass Appeal: Metabolite identification in mass spectrometry-focused untargeted metabolomics. Metabolomics 2013, 9, S44–S66. [Google Scholar]
  694. Möller, M.; Du Preez, J.L.; Harvey, B.H. Development and validation of a single analytical method for the determination of tryptophan, and its kynurenine metabolites in rat plasma. J. Chromatogr. B Analyt Technol. Biomed. Life Sci. 2012, 898, 121–129. [Google Scholar] [CrossRef] [PubMed]
  695. Qiu, S.; Cai, Y.; Yao, H.; Lin, C.; Xie, Y.; Tang, S.; Zhang, A. Small molecule metabolites: Discovery of biomarkers and therapeutic targets. Signal Transduct. Target. Ther. 2023, 8, 132. [Google Scholar] [CrossRef]
  696. Patel, V.D.; Shamsi, S.A.; Miller, A.; Liu, A.; Powell, M. Simultaneous separation and detection of nine kynurenine pathway metabolites by reversed-phase liquid chromatography-mass spectrometry: Quantitation of inflammation in human cerebrospinal fluid and plasma. Anal. Chim. Acta 2023, 1278, 341659. [Google Scholar] [CrossRef] [PubMed]
  697. Abusoglu, S.; Eryavuz Onmaz, D.; Abusoglu, G.; Humeyra Yerlikaya, F.; Unlu, A. Measurement of kynurenine pathway metabolites by tandem mass spectrometry. J. Mass. Spectrom. Adv. Clin. Lab. 2023, 28, 114–121. [Google Scholar] [CrossRef] [PubMed]
  698. Fuertig, R.; Ceci, A.; Camus, S.M.; Bezard, E.; Luippold, A.H.; Hengerer, B. LC-MS/MS-based quantification of kynurenine metabolites, tryptophan, monoamines and neopterin in plasma, cerebrospinal fluid and brain. Bioanalysis 2016, 8, 1903–1917. [Google Scholar] [CrossRef] [PubMed]
  699. Heng, B.; Pires, A.S.; Chow, S.; Krishnamurthy, S.; Bonnell, B.; Bustamante, S.; Guillemin, G.J. Stability Studies of Kynurenine Pathway Metabolites in Blood Components Define Optimal Blood Processing Conditions. Int. J. Tryptophan Res. 2023, 16, 11786469231213521. [Google Scholar] [CrossRef] [PubMed]
  700. Fukushima, T.; Umino, M.; Sakamoto, T.; Onozato, M. A review of chromatographic methods for bioactive tryptophan metabolites, kynurenine, kynurenic acid, quinolinic acid, and others, in biological fluids. Biomed. Chromatogr. 2022, 36, e5308. [Google Scholar] [CrossRef]
  701. Mitsuhashi, S.; Fukushima, T.; Arai, K.; Tomiya, M.; Santa, T.; Imai, K.; Toyo’oka, T. Development of a column-switching high-performance liquid chromatography for kynurenine enantiomers and its application to a pharmacokinetic study in rat plasma. Anal. Chim. Acta 2007, 587, 60–66. [Google Scholar] [CrossRef] [PubMed]
  702. Pi, L.G.; Tang, A.G.; Mo, X.M.; Luo, X.B.; Ao, X. More rapid and sensitive method for simultaneous determination of tryptophan and kynurenic acid by HPLC. Clin. Biochem. 2009, 42, 420–425. [Google Scholar] [CrossRef] [PubMed]
  703. Bao, Y.; Luchetti, D.; Schaeffer, E.; Cutrone, J. Determination of kynurenic acid in rat cerebrospinal fluid by HPLC with fluorescence detection. Biomed. Chromatogr. 2016, 30, 62–67. [Google Scholar] [CrossRef] [PubMed]
  704. Soto, M.E.; Ares, A.M.; Bernal, J.; Nozal, M.J.; Bernal, J.L. Simultaneous determination of tryptophan, kynurenine, kynurenic and xanthurenic acids in honey by liquid chromatography with diode array, fluorescence and tandem mass spectrometry detection. J. Chromatogr. A 2011, 1218, 7592–7600. [Google Scholar] [CrossRef] [PubMed]
  705. Sadok, I.; Gamian, A.; Staniszewska, M.M. Chromatographic analysis of tryptophan metabolites. J. Sep. Sci. 2017, 40, 3020–3045. [Google Scholar] [CrossRef] [PubMed]
  706. Marković, M.; Petronijević, N.; Stašević, M.; Stašević Karličić, I.; Velimirović, M.; Stojković, T.; Ristić, S.; Stojković, M.; Milić, N.; Nikolić, T. Decreased Plasma Levels of Kynurenine and Kynurenic Acid in Previously Treated and First-Episode Antipsychotic-Naive Schizophrenia Patients. Cells 2023, 12, 2814. [Google Scholar] [CrossRef] [PubMed]
  707. Mawatari, K.; Iinuma, F.; Watanabe, M. Fluorometric determination of urinary kynurenic acid by flow injection analysis equipped with a “bypass line”. Anal. Biochem. 1990, 190, 88–91. [Google Scholar] [CrossRef] [PubMed]
  708. Atsumi, M.; Mawatari, K.I.; Morooka, A.; Yasuda, M.; Fukuuchi, T.; Yamaoka, N.; Kaneko, K.; Nakagomi, K.; Oku, N. Simultaneous Determination of Kynurenine and Kynurenic Acid by High-Performance Liquid Chromatography Photoirradiation System Using a Mobile Phase Containing 18-crown-6. Int. J. Tryptophan Res. 2019, 12, 1178646919834551. [Google Scholar] [CrossRef] [PubMed]
  709. Odo, J.; Funai, T.; Hirai, A. Spectrofluorometric determination of kynurenic acid with horseradish peroxidase in the presence of hydrogen peroxide. Anal. Sci. 2007, 23, 317–320. [Google Scholar] [CrossRef] [PubMed]
  710. Odo, J.; Inoguchi, M.; Aoki, H.; Sogawa, Y.; Nishimura, M. Fluorescent derivatization of aromatic carboxylic acids with horseradish peroxidase in the presence of excess hydrogen peroxide. Anal. Sci. 2015, 31, 37–44. [Google Scholar] [CrossRef] [PubMed]
  711. Lu, Y.; Lin, L.; Ye, J. Human metabolite detection by surface-enhanced Raman spectroscopy. Mater. Today Bio 2022, 13, 100205. [Google Scholar] [CrossRef]
  712. Murphy, M.P.; O’Neill, L.A.J. A break in mitochondrial endosymbiosis as a basis for inflammatory diseases. Nature 2024, 626, 271–279. [Google Scholar] [CrossRef]
  713. Mangoni, A.A.; Zinellu, A. A systematic review and meta-analysis of the kynurenine pathway of tryptophan metabolism in rheumatic diseases. Front. Immunol. 2023, 14, 1257159. [Google Scholar] [CrossRef]
  714. Encarnacion, A.B.; Fagutao, F.; Hirono, I.; Ushio, H.; Ohshima, T. Effects of ergothioneine from mushrooms (Flammulina velutipes) on melanosis and lipid oxidation of kuruma shrimp (Marsupenaeus japonicus). J. Agric. Food Chem. 2010, 58, 2577–2585. [Google Scholar] [CrossRef] [PubMed]
  715. Encarnacion, A.B.; Fagutao, F.; Hirayama, J.; Terayama, M.; Hirono, I.; Ohshima, T. Edible mushroom (Flammulina velutipes) extract inhibits melanosis in Kuruma shrimp (Marsupenaeus japonicus). J. Food Sci. 2011, 76, C52-58. [Google Scholar] [CrossRef] [PubMed]
  716. Encarnacion, A.B.; Fagutao, F.; Shozen, K.; Hirono, I.; Ohshima, T. Biochemical intervention of ergothioneine-rich edible mushroom (Flammulina velutipes) extract inhibits melanosis in crab (Chionoecetes japonicus). Food Chem. 2011, 127, 1594–1599. [Google Scholar] [CrossRef]
  717. Staniszewska, M.; Kowalik, S.; Sadok, I.; Kędzierski, W. The Influence of Exercise Intensity on Tryptophan Metabolites in Thoroughbred Horses. Pharmaceuticals 2023, 16, 107. [Google Scholar] [CrossRef]
  718. Mago, Y.; Sharma, Y.; Thakran, Y.; Mishra, A.; Tewari, S.; Kataria, N. Next-Generation Organic Beauty Products Obtained from Algal Secondary Metabolites: A Sustainable Development in Cosmeceutical Industries. Mol. Biotechnol. 2023. [Google Scholar] [CrossRef]
  719. Goyal, A.; Sharma, A.; Kaur, J.; Kumari, S.; Garg, M.; Sindhu, R.K.; Rahman, M.H.; Akhtar, M.F.; Tagde, P.; Najda, A.; et al. Bioactive-Based Cosmeceuticals: An Update on Emerging Trends. Molecules 2022, 27, 828. [Google Scholar] [CrossRef]
  720. Vaishampayan, P.; Rane, M.M. Herbal nanocosmecuticals: A review on cosmeceutical innovation. J. Cosmet. Dermatol. 2022, 21, 5464–5483. [Google Scholar] [CrossRef]
  721. Lee, C.M. Fifty years of research and development of cosmeceuticals: A contemporary review. J. Cosmet. Dermatol. 2016, 15, 527–539. [Google Scholar] [CrossRef] [PubMed]
  722. Epstein, H. Cosmeceuticals and polyphenols. Clin. Dermatol. 2009, 27, 475–478. [Google Scholar] [CrossRef] [PubMed]
  723. Taofiq, O.; González-Paramás, A.M.; Martins, A.; Barreiro, M.F.; Ferreira, I.C.F.R. Mushrooms extracts and compounds in cosmetics, cosmeceuticals and nutricosmetics—A review. Ind. Crops Prod. 2016, 90, 38–48. [Google Scholar] [CrossRef]
  724. Pérez-Sánchez, A.; Barrajón-Catalán, E.; Herranz-López, M.; Micol, V. Nutraceuticals for Skin Care: A Comprehensive Review of Human Clinical Studies. Nutrients 2018, 10, 403. [Google Scholar] [CrossRef] [PubMed]
  725. Norins, A.L. Free radical formation in the skin following exposure to ultraviolet light. J. Investig. Dermatol. 1962, 39, 445–448. [Google Scholar]
  726. Chen, X.; Yang, C.; Jiang, G. Research progress on skin photoaging and oxidative stress. Postepy Dermatol. Alergol. 2021, 38, 931–936. [Google Scholar] [CrossRef]
  727. Fernandes, A.; Rodrigues, P.M.; Pintado, M.; Tavaria, F.K. A systematic review of natural products for skin applications: Targeting inflammation, wound healing, and photo-aging. Phytomedicine 2023, 115, 154824. [Google Scholar] [CrossRef]
  728. Salminen, A.; Kaarniranta, K.; Kauppinen, A. Photoaging: UV radiation-induced inflammation and immunosuppression accelerate the aging process in the skin. Inflamm. Res. 2022, 71, 817–831. [Google Scholar] [CrossRef]
  729. Bissonnette, R.; Stein Gold, L.; Rubenstein, D.S.; Tallman, A.M.; Armstrong, A. Tapinarof in the treatment of psoriasis: A review of the unique mechanism of action of a novel therapeutic aryl hydrocarbon receptor-modulating agent. J. Am. Acad. Dermatol. 2021, 84, 1059–1067. [Google Scholar] [CrossRef]
  730. Bissonnette, R.; Saint-Cyr Proulx, E.; Jack, C.; Maari, C. Tapinarof for psoriasis and atopic dermatitis: 15 years of clinical research. J. Eur. Acad. Dermatol. Venereol. 2023, 37, 1168–1174. [Google Scholar] [CrossRef]
  731. Furue, M.; Hashimoto-Hachiya, A.; Tsuji, G. Aryl Hydrocarbon Receptor in Atopic Dermatitis and Psoriasis. Int. J. Mol. Sci. 2019, 20, 5424. [Google Scholar] [CrossRef]
  732. Tsuji, G.; Hashimoto-Hachiya, A.; Matsuda-Taniguchi, T.; Takai-Yumine, A.; Takemura, M.; Yan, X.; Furue, M.; Nakahara, T. Natural Compounds Tapinarof and Galactomyces Ferment Filtrate Downregulate IL-33 Expression via the AHR/IL-37 Axis in Human Keratinocytes. Front. Immunol. 2022, 13, 745997. [Google Scholar] [CrossRef]
  733. Hwang, J.; Newton, E.M.; Hsiao, J.; Shi, V.Y. Aryl hydrocarbon receptor/nuclear factor E2-related factor 2 (AHR/NRF2) signalling: A novel therapeutic target for atopic dermatitis. Exp. Dermatol. 2022, 31, 485–497. [Google Scholar] [CrossRef] [PubMed]
  734. Van den Bogaard, E.H.; Bergboer, J.G.M.; Vonk-Bergers, M.; van Vlijmen-Willems, I.M.J.J.; Hato, S.V.; van der Valk, P.G.M.; Schröder, J.M.; Joosten, I.; Zeeuwen, P.L.J.M.; Schalkwijk, J. Coal tar induces AHR-dependent skin barrier repair in atopic dermatitis. J. Clin. Investig. 2013, 123, 917–927. [Google Scholar] [CrossRef] [PubMed]
  735. Lee, J.; Song, K.M.; Jung, C.H. Diosmin restores the skin barrier by targeting the aryl hydrocarbon receptor in atopic dermatitis. Phytomedicine 2021, 81, 153418. [Google Scholar] [CrossRef]
  736. Li, X.; Gianoulis, T.A.; Yip, K.Y.; Gerstein, M.; Snyder, M. Extensive in vivo metabolite-protein interactions revealed by large-scale systematic analyses. Cell 2010, 143, 639–650. [Google Scholar]
  737. Li, X.; Snyder, M. Metabolites as global regulators: A new view of protein regulation: Systematic investigation of metabolite-protein interactions may help bridge the gap between genome-wide association studies and small molecule screening studies. Bioessays 2011, 33, 485–489. [Google Scholar] [PubMed]
  738. Kell, D.B. Metabolites do social networking. Nat. Chem. Biol. 2011, 7, 7–8. [Google Scholar]
  739. Piazza, I.; Kochanowski, K.; Cappelletti, V.; Fuhrer, T.; Noor, E.; Sauer, U.; Picotti, P. A map of protein-metabolite interactions reveals principles of chemical communication. Cell 2018, 172, 358–372.e323. [Google Scholar] [CrossRef]
  740. Malinovska, L.; Cappelletti, V.; Kohler, D.; Piazza, I.; Tsai, T.H.; Pepelnjak, M.; Stalder, P.; Dorig, C.; Sesterhenn, F.; Elsasser, F.; et al. Proteome-wide structural changes measured with limited proteolysis-mass spectrometry: An advanced protocol for high-throughput applications. Nat. Protoc. 2023, 18, 659–682. [Google Scholar] [CrossRef]
  741. Hseu, Y.C.; Lo, H.W.; Korivi, M.; Tsai, Y.C.; Tang, M.J.; Yang, H.L. Dermato-protective properties of ergothioneine through induction of Nrf2/ARE-mediated antioxidant genes in UVA-irradiated Human keratinocytes. Free Radic. Biol. Med. 2015, 86, 102–117. [Google Scholar] [CrossRef] [PubMed]
  742. Zalachoras, I.; Hollis, F.; Ramos-Fernández, E.; Trovo, L.; Sonnay, S.; Geiser, E.; Preitner, N.; Steiner, P.; Sandi, C.; Morató, L. Therapeutic potential of glutathione-enhancers in stress-related psychopathologies. Neurosci. Biobehav. Rev. 2020, 114, 134–155. [Google Scholar] [CrossRef]
  743. Hseu, Y.C.; Vudhya Gowrisankar, Y.; Chen, X.Z.; Yang, Y.C.; Yang, H.L. The Antiaging Activity of Ergothioneine in UVA-Irradiated Human Dermal Fibroblasts via the Inhibition of the AP-1 Pathway and the Activation of Nrf2-Mediated Antioxidant Genes. Oxid. Med. Cell Longev. 2020, 2020, 2576823. [Google Scholar] [CrossRef]
  744. Bernardo, V.S.; Torres, F.F.; de Paula, C.P.; de Oliveira da Silva, J.P.M.; de Almeida, E.A.; da Cunha, A.F.; da Silva, D.G.H. Potential Cytoprotective and Regulatory Effects of Ergothioneine on Gene Expression of Proteins Involved in Erythroid Adaptation Mechanisms and Redox Pathways in K562 Cells. Genes 2022, 13, 2368. [Google Scholar] [CrossRef] [PubMed]
  745. Dare, A.; Elrashedy, A.A.; Channa, M.L.; Nadar, A. Cardioprotective Effects and in-silico Antioxidant Mechanism of L-Ergothioneine in Experimental Type-2 Diabetic Rats. Cardiovasc. Hematol. Agents Med. Chem. 2022, 20, 133–147. [Google Scholar] [CrossRef]
  746. Jomova, K.; Raptova, R.; Alomar, S.Y.; Alwasel, S.H.; Nepovimova, E.; Kuca, K.; Valko, M. Reactive oxygen species, toxicity, oxidative stress, and antioxidants: Chronic diseases and aging. Arch. Toxicol. 2023, 97, 2499–2574. [Google Scholar] [CrossRef]
  747. Kim, H.-J.; Jin, B.-R.; Lee, C.-D.; Kim, D.; Lee, A.Y.; Lee, S.; An, H.-J. Anti-Inflammatory Effect of Chestnut Honey and Cabbage Mixtures Alleviates Gastric Mucosal Damage. Nutrients 2024, 16, 389. [Google Scholar] [CrossRef]
  748. Pierozan, P.; Biasibetti-Brendler, H.; Schmitz, F.; Ferreira, F.; Pessoa-Pureur, R.; Wyse, A.T.S. Kynurenic Acid Prevents Cytoskeletal Disorganization Induced by Quinolinic Acid in Mixed Cultures of Rat Striatum. Mol. Neurobiol. 2018, 55, 5111–5124. [Google Scholar] [CrossRef]
  749. Yeager, R.L.; Reisman, S.A.; Aleksunes, L.M.; Klaassen, C.D. Introducing the “TCDD-inducible AhR-Nrf2 gene battery”. Toxicol. Sci. 2009, 111, 238–246. [Google Scholar] [CrossRef] [PubMed]
  750. Takei, K.; Mitoma, C.; Hashimoto-Hachiya, A.; Uchi, H.; Takahara, M.; Tsuji, G.; Kido-Nakahara, M.; Nakahara, T.; Furue, M. Antioxidant soybean tar Glyteer rescues T-helper-mediated downregulation of filaggrin expression via aryl hydrocarbon receptor. J. Dermatol. 2015, 42, 171–180. [Google Scholar] [CrossRef]
  751. Furue, M.; Ishii, Y.; Tsukimori, K.; Tsuji, G. Aryl Hydrocarbon Receptor and Dioxin-Related Health Hazards-Lessons from Yusho. Int. J. Mol. Sci. 2021, 22, 708. [Google Scholar] [CrossRef]
  752. Murray, I.A.; Patterson, A.D.; Perdew, G.H. Aryl hydrocarbon receptor ligands in cancer: Friend and foe. Nat. Rev. Cancer 2014, 14, 801–814. [Google Scholar] [CrossRef]
  753. Vogel, C.F.; Khan, E.M.; Leung, P.S.; Gershwin, M.E.; Chang, W.L.; Wu, D.; Haarmann-Stemmann, T.; Hoffmann, A.; Denison, M.S. Cross-talk between aryl hydrocarbon receptor and the inflammatory response: A role for nuclear factor-κB. J. Biol. Chem. 2014, 289, 1866–1875. [Google Scholar] [CrossRef] [PubMed]
  754. Nelson, D.E.; Ihekwaba, A.E.C.; Elliott, M.; Gibney, C.A.; Foreman, B.E.; Nelson, G.; See, V.; Horton, C.A.; Spiller, D.G.; Edwards, S.W.; et al. Oscillations in NF-κB signalling control the dynamics of gene expression. Science 2004, 306, 704–708. [Google Scholar]
  755. Kell, D.B. Metabolomics, modelling and machine learning in systems biology: Towards an understanding of the languages of cells. The 2005 Theodor Bücher lecture. FEBS J 2006, 273, 873–894. [Google Scholar]
  756. Ashall, L.; Horton, C.A.; Nelson, D.E.; Paszek, P.; Ryan, S.; Sillitoe, K.; Harper, C.V.; Spiller, D.G.; Unitt, J.F.; Broomhead, D.S.; et al. Pulsatile stimulation determines timing and specificity of NFkappa-B-dependent transcription. Science 2009, 324, 242–246. [Google Scholar] [PubMed]
  757. Bento, A.P.; Hersey, A.; Félix, E.; Landrum, G.; Gaulton, A.; Atkinson, F.; Bellis, L.J.; De Veij, M.; Leach, A.R. An open source chemical structure curation pipeline using RDKit. J. Cheminform 2020, 12, 51. [Google Scholar] [CrossRef] [PubMed]
  758. O’Hagan, S.; Swainston, N.; Handl, J.; Kell, D.B. A ‘rule of 0.5’ for the metabolite-likeness of approved pharmaceutical drugs. Metabolomics 2015, 11, 323–339. [Google Scholar] [CrossRef]
  759. O’Hagan, S.; Kell, D.B. Understanding the foundations of the structural similarities between marketed drugs and endogenous human metabolites. Front. Pharmacol. 2015, 6, 105. [Google Scholar] [CrossRef]
  760. O’Hagan, S.; Kell, D.B. The KNIME workflow environment and its applications in Genetic Programming and machine learning. Genetic Progr Evol. Mach. 2015, 16, 387–391. [Google Scholar] [CrossRef]
  761. O’Hagan, S.; Kell, D.B. Consensus rank orderings of molecular fingerprints illustrate the ‘most genuine’ similarities between marketed drugs and small endogenous human metabolites, but highlight exogenous natural products as the most important ‘natural’ drug transporter substrates. ADMET & DMPK 2017, 5, 85–125. [Google Scholar] [CrossRef]
  762. Lin, X.M.; Li, H.; Wang, C.; Peng, X.X. Proteomic analysis of nalidixic acid resistance in Escherichia coli: Identification and functional characterization of OM proteins. J. Proteome Res. 2008, 7, 2399–2405. [Google Scholar] [CrossRef] [PubMed]
  763. Munro, L.J.; Kell, D.B. Analysis of a library of E. coli transporter knockout strains to identify transport pathways of antibiotics. Antibiotics 2022, 11, 1129. [Google Scholar] [CrossRef]
  764. Ren, X.; Yan, C.X.; Zhai, R.X.; Xu, K.; Li, H.; Fu, X.J. Comprehensive survey of target prediction web servers for Traditional Chinese Medicine. Heliyon 2023, 9, e19151. [Google Scholar] [CrossRef] [PubMed]
  765. Hagg, A.; Kirschner, K.N. Open-Source Machine Learning in Computational Chemistry. J. Chem. Inf. Model. 2023, 63, 4505–4532. [Google Scholar] [CrossRef]
  766. Kell, D.B.; Samanta, S.; Swainston, N. Deep learning and generative methods in cheminformatics and chemical biology: Navigating small molecule space intelligently. Biochem. J. 2020, 477, 4559–4580. [Google Scholar] [CrossRef]
  767. Wang, X.; Shen, Y.; Wang, S.; Li, S.; Zhang, W.; Liu, X.; Lai, L.; Pei, J.; Li, H. PharmMapper 2017 update: A web server for potential drug target identification with a comprehensive target pharmacophore database. Nucleic Acids Res. 2017, 45, W356–W360. [Google Scholar] [CrossRef]
  768. Daina, A.; Michielin, O.; Zoete, V. SwissTargetPrediction: Updated data and new features for efficient prediction of protein targets of small molecules. Nucleic Acids Res. 2019, 47, W357–W364. [Google Scholar] [CrossRef]
  769. Gallo, K.; Goede, A.; Preissner, R.; Gohlke, B.O. SuperPred 3.0: Drug classification and target prediction—A machine learning approach. Nucleic Acids Res. 2022, 50, W726–W731. [Google Scholar] [CrossRef]
  770. Weininger, D. SMILES, a chemical language and information system. 1. Introduction to methodology and encoding rules. J. Chem. Inf. Comput. Sci. 1988, 28, 31–36. [Google Scholar]
  771. Temperini, C.; Innocenti, A.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators: Kinetic and X-ray crystallographic study for the interaction of D- and L-tryptophan with the mammalian isoforms I-XIV. Bioorg Med. Chem. 2008, 16, 8373–8378. [Google Scholar] [CrossRef] [PubMed]
  772. Moore, J.B.; Blanchard, R.K.; McCormack, W.T.; Cousins, R.J. cDNA array analysis identifies thymic LCK as upregulated in moderate murine zinc deficiency before T-lymphocyte population changes. J. Nutr. 2001, 131, 3189–3196. [Google Scholar] [CrossRef] [PubMed]
  773. Sherman, B.T.; Hao, M.; Qiu, J.; Jiao, X.; Baseler, M.W.; Lane, H.C.; Imamichi, T.; Chang, W. DAVID: A web server for functional enrichment analysis and functional annotation of gene lists (2021 update). Nucleic Acids Res. 2022, 50, W216–W221. [Google Scholar] [CrossRef]
  774. Lörinczi, B.; Csámpai, A.; Fülöp, F.; Szatmári, I. Synthesis of New C-3 Substituted Kynurenic Acid Derivatives. Molecules 2020, 25, 937. [Google Scholar] [CrossRef]
  775. Simon, P.; Lörinczi, B.; Hetényi, A.; Szatmári, I. Novel Eco-friendly, One-Pot Method for the Synthesis of Kynurenic Acid Ethyl Esters. ACS Omega 2023, 8, 17966–17975. [Google Scholar] [CrossRef] [PubMed]
  776. Verma, S.; Ambatwar, R.; Datusalia, A.K.; Khatik, G.L. Convenient One-Pot Synthesis of Kynurenic Acid Ethyl Ester and Exploration to Direct Synthesis of Neuroprotective Kynurenic Acid and Amide Derivatives. J. Org. Chem. 2023, 88, 10494–10500. [Google Scholar] [CrossRef] [PubMed]
  777. van der Hoek, S.A.; Darbani, B.; Zugaj, K.; Prabhala, B.K.; Biron, M.B.; Randelovic, M.; Medina, J.B.; Kell, D.B.; Borodina, I. Engineering the yeast Saccharomyces cerevisiae for the production of L-(+)-ergothioneine. Front. Bioeng. Biotechnol. 2019, 7, 262. [Google Scholar] [CrossRef]
  778. Van der Hoek, S.A.; Rusnák, M.; Wang, G.; Stanchev, L.D.; de Fátima Alvez, L.; Jessop-Fabre, M.M.; Paramasivan, K.; Jacobsen, I.H.; Sonnenschein, N.; Martínez, J.L.; et al. Engineering precursor supply for the high-level production of ergothioneine in Saccharomyces cerevisiae. Metab. Eng. 2022, 70, 129–142. [Google Scholar]
  779. van der Hoek, S.A.; Rusnák, M.; Jacobsen, I.H.; Martínez, J.L.; Kell, D.B.; Borodina, I. Engineering ergothioneine production in Yarrowia lipolytica. FEBS Lett. 2022, 596, 1356–1364. [Google Scholar] [CrossRef]
  780. Sharma, K.; Ghiffary, M.R.; Lee, G.; Kim, H.U. Efficient production of an antitumor precursor actinocin and other medicinal molecules from kynurenine pathway in Escherichia coli. Metab. Eng. 2023, 81, 144–156. [Google Scholar] [CrossRef]
  781. Dei Cas, M.; Vigentini, I.; Vitalini, S.; Laganaro, A.; Iriti, M.; Paroni, R.; Foschino, R. Tryptophan Derivatives by Saccharomyces cerevisiae EC1118: Evaluation, Optimization, and Production in a Soybean-Based Medium. Int. J. Mol. Sci. 2021, 22, 472. [Google Scholar] [CrossRef] [PubMed]
  782. Yılmaz, C.; Gökmen, V. Kinetic evaluation of the formation of tryptophan derivatives in the kynurenine pathway during wort fermentation using Saccharomyces pastorianus and Saccharomyces cerevisiae. Food Chem. 2019, 297, 124975. [Google Scholar] [CrossRef]
  783. Wróbel-Kwiatkowska, M.; Turski, W.; Kocki, T.; Rakicka-Pustułka, M.; Rymowicz, W. An efficient method for production of kynurenic acid by Yarrowia lipolytica. Yeast 2020, 37, 541–547. [Google Scholar] [CrossRef]
  784. Rakicka-Pustułka, M.; Ziuzia, P.; Pierwola, J.; Szymanski, K.; Wróbel-Kwiatkowska, M.; Lazar, Z. The microbial production of kynurenic acid using Yarrowia lipolytica yeast growing on crude glycerol and soybean molasses. Front. Bioeng. Biotechnol. 2022, 10, 936137. [Google Scholar] [CrossRef]
  785. Matusiewicz, M.; Wróbel-Kwiatkowska, M.; Niemiec, T.; Świderek, W.; Kosieradzka, I.; Rosińska, A.; Niwińska, A.; Rakicka-Pustułka, M.; Kocki, T.; Rymowicz, W.; et al. Effect of Yarrowia lipolytica yeast biomass with increased kynurenic acid content on selected metabolic indicators in mice. PeerJ 2023, 11, e15833. [Google Scholar] [CrossRef]
  786. Aburto, J.M.; Villavicencio, F.; Basellini, U.; Kjaergaard, S.; Vaupel, J.W. Dynamics of life expectancy and life span equality. Proc. Natl. Acad. Sci. USA 2020, 117, 5250–5259. [Google Scholar] [CrossRef]
  787. Crimmins, E.M. Lifespan and Healthspan: Past, Present, and Promise. Gerontologist 2015, 55, 901–911. [Google Scholar] [CrossRef] [PubMed]
  788. Akbaraly, T.N.; Singh-Manoux, A.; Tabak, A.G.; Jokela, M.; Virtanen, M.; Ferrie, J.E.; Marmot, M.G.; Shipley, M.J.; Kivimaki, M. Overall diet history and reversibility of the metabolic syndrome over 5 years: The Whitehall II prospective cohort study. Diabetes Care 2010, 33, 2339–2341. [Google Scholar] [CrossRef]
  789. Lagström, H.; Stenholm, S.; Akbaraly, T.; Pentti, J.; Vahtera, J.; Kivimäki, M.; Head, J. Diet quality as a predictor of cardiometabolic disease-free life expectancy: The Whitehall II cohort study. Am. J. Clin. Nutr. 2020, 111, 787–794. [Google Scholar] [CrossRef] [PubMed]
  790. Soda, K. Spermine and gene methylation: A mechanism of lifespan extension induced by polyamine-rich diet. Amino Acids 2020, 52, 213–224. [Google Scholar] [CrossRef]
  791. Smith, E.; Ottosson, F.; Hellstrand, S.; Ericson, U.; Orho-Melander, M.; Fernandez, C.; Melander, O. Ergothioneine is associated with reduced mortality and decreased risk of cardiovascular disease. Heart 2020, 106, 691–697. [Google Scholar] [CrossRef] [PubMed]
  792. Halliwell, B.; Tang, R.M.Y.; Cheah, I.K. Diet-Derived Antioxidants: The Special Case of Ergothioneine. Annu. Rev. Food Sci. Technol. 2023, 14, 323–345. [Google Scholar] [CrossRef]
  793. Tian, X.; Cioccoloni, G.; Sier, J.H.; Naseem, K.M.; Thorne, J.L.; Moore, J.B. Ergothioneine supplementation in people with metabolic syndrome (ErgMS): Protocol for a randomised, double-blind, placebocontrolled pilot study. Pilot. Feas Stud. 2021, 7, 193. [Google Scholar] [CrossRef]
  794. Cheah, I.K.; Halliwell, B. Ergothioneine, recent developments. Redox Biol. 2021, 42, 101868. [Google Scholar] [CrossRef]
  795. Beelman, R.B.; Phillips, A.T.; Richie, J.P., Jr.; Ba, D.M.; Duiker, S.W.; Kalaras, M.D. Health Consequences of Improving the Content of Ergothioneine in the Food Supply. FEBS Lett. 2022, 596, 1231–1240. [Google Scholar] [CrossRef]
  796. Ey, J.; Schömig, E.; Taubert, D. Dietary sources and antioxidant effects of ergothioneine. J. Agric. Food Chem. 2007, 55, 6466–6474. [Google Scholar] [CrossRef]
  797. Kalaras, M.D.; Richie, J.P.; Calcagnotto, A.; Beelman, R.B. Mushrooms: A rich source of the antioxidants ergothioneine and glutathione. Food Chem. 2017, 233, 429–433. [Google Scholar] [CrossRef]
  798. Agrawal, D.C.; Dhanasekaran, M. (Eds.) Medicinal Mushrooms: Recent Progress in Research and Development; Springer: Singapore, 2019. [Google Scholar]
  799. Uffelman, C.N.; Doenges, K.A.; Armstrong, M.L.; Quinn, K.; Reisdorph, R.M.; Tang, M.; Krebs, N.F.; Reisdorph, N.A.; Campbell, W.W. Metabolomics Profiling of White Button, Crimini, Portabella, Lion’s Mane, Maitake, Oyster, and Shiitake Mushrooms Using Untargeted Metabolomics and Targeted Amino Acid Analysis. Foods 2023, 12, 2985. [Google Scholar] [CrossRef]
  800. Tsiantas, K.; Tsiaka, T.; Koutrotsios, G.; Siapi, E.; Zervakis, G.I.; Kalogeropoulos, N.; Zoumpoulakis, P. On the Identification and Quantification of Ergothioneine and Lovastatin in Various Mushroom Species: Assets and Challenges of Different Analytical Approaches. Molecules 2021, 26, 1832. [Google Scholar] [CrossRef]
  801. Mitsuyama, H.; May, J.M. Uptake and antioxidant effects of ergothioneine in human erythrocytes. Clin. Sci. 1999, 97, 407–411. [Google Scholar]
  802. Gründemann, D.; Harlfinger, S.; Golz, S.; Geerts, A.; Lazar, A.; Berkels, R.; Jung, N.; Rubbert, A.; Schömig, E. Discovery of the ergothioneine transporter. Proc. Natl. Acad. Sci. USA 2005, 102, 5256–5261. [Google Scholar] [CrossRef] [PubMed]
  803. Gründemann, D.; Hartmann, L.; Flögel, S. The Ergothioneine Transporter (ETT): Substrates and Locations, an Inventory. FEBS Lett. 2022, 596, 1252–1269. [Google Scholar] [CrossRef] [PubMed]
  804. Yee, S.W.; Buitrago, D.; Stecula, A.; Ngo, H.X.; Chien, H.C.; Zou, L.; Koleske, M.L.; Giacomini, K.M. Deorphaning a solute carrier 22 family member, SLC22A15, through functional genomic studies. FASEB J. 2020, 34, 15734–15752. [Google Scholar] [CrossRef]
  805. Wang, L.Z.; Thuya, W.L.; Toh, D.S.; Lie, M.G.; Lau, J.Y.; Kong, L.R.; Wan, S.C.; Chua, K.N.; Lee, E.J.; Goh, B.C. Quantification of L-ergothioneine in human plasma and erythrocytes by liquid chromatography-tandem mass spectrometry. J. Mass. Spectrom. 2013, 48, 406–412. [Google Scholar] [CrossRef] [PubMed]
  806. Taubert, D.; Lazar, A.; Grimberg, G.; Jung, N.; Rubbert, A.; Delank, K.S.; Perniok, A.; Erdmann, E.; Schömig, E. Association of rheumatoid arthritis with ergothioneine levels in red blood cells: A case control study. J. Rheumatol. 2006, 33, 2139–2145. [Google Scholar]
  807. Flegel, W.A.; Srivastava, K.; Sissung, T.M.; Goldspiel, B.R.; Figg, W.D. Pharmacogenomics with red cells: A model to study protein variants of drug transporter genes. Vox Sang. 2021, 116, 141–154. [Google Scholar] [CrossRef]
  808. Zhang, Y.; Gonzalez-Gutierrez, G.; Legg, K.A.; Walsh, B.J.C.; Pis Diez, C.M.; Edmonds, K.A.; Giedroc, D.P. Discovery and structure of a widespread bacterial ABC transporter specific for ergothioneine. Nat. Commun. 2022, 13, 7586. [Google Scholar] [CrossRef]
  809. Dumitrescu, D.G.; Gordon, E.M.; Kovalyova, Y.; Seminara, A.B.; Duncan-Lowey, B.; Forster, E.R.; Zhou, W.; Booth, C.J.; Shen, A.; Kranzusch, P.J.; et al. A microbial transporter of the dietary antioxidant ergothioneine. Cell 2022, 185, 4526–4540.e18. [Google Scholar] [CrossRef]
  810. Alamgir, K.M.; Masuda, S.; Fujitani, Y.; Fukuda, F.; Tani, A. Production of ergothioneine by Methylobacterium species. Front. Microbiol. 2015, 6, 1185. [Google Scholar] [CrossRef]
  811. Pepelnjak, M.; de Souza, N.; Picotti, P. Detecting Protein-Small Molecule Interactions Using Limited Proteolysis-Mass Spectrometry (LiP-MS). Trends Biochem. Sci. 2020, 45, 919–920. [Google Scholar] [CrossRef]
  812. Piazza, I.; Beaton, N.; Bruderer, R.; Knobloch, T.; Barbisan, C.; Chandat, L.; Sudau, A.; Siepe, I.; Rinner, O.; de Souza, N.; et al. A machine learning-based chemoproteomic approach to identify drug targets and binding sites in complex proteomes. Nat. Commun. 2020, 11, 4200. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Kynurenic acid structure and tautomers.
Figure 1. Kynurenic acid structure and tautomers.
Ijms 25 09082 g001
Figure 2. Elements of the kynurenine pathway. While the cytoprotective kynurenic acid can be derived endogenously from tryptophan, the pathway involves the synthesis of kynurenine that can lead to other toxic products such as quinolinic acid and 3-hydroxykynurenine. Redrawn in part from [102]. * Indicates ring cyclisation.
Figure 2. Elements of the kynurenine pathway. While the cytoprotective kynurenic acid can be derived endogenously from tryptophan, the pathway involves the synthesis of kynurenine that can lead to other toxic products such as quinolinic acid and 3-hydroxykynurenine. Redrawn in part from [102]. * Indicates ring cyclisation.
Ijms 25 09082 g002
Figure 3. The ‘terminal’ steps of the KP from tryptophan to kynurenic acid. 1. L-tryptophan is catalyzed to N-formyl-L-kynurenine (NFK) by tryptophan dioxygenase (TDO) or indole dioxygenase (IDO1, IDO2) (EC 1.13.11.11 and 1.13.11.52), depending on the organism/tissue. 2. NFK is then converted to L-kynurenine (KYN) by the kynurenine formamidase (E.C. 3.5.1.9). 3. Finally, KYN is catalyzed to the unstable 4-(2-aminophenyl)-2,4-dioxobutanoate intermediate by a kynurenine transaminase (KAT1-4) (E. C. 2.6.1.7), which is readily converted to KYNA by a spontaneous reaction. The spontaneous cyclization of the intermediate to KYNA is unique to KYNA biosynthesis, and it makes this reaction effectively irreversible meaning that exogenous KYNA will not be converted to L-kynurenine nor its toxic derivatives.
Figure 3. The ‘terminal’ steps of the KP from tryptophan to kynurenic acid. 1. L-tryptophan is catalyzed to N-formyl-L-kynurenine (NFK) by tryptophan dioxygenase (TDO) or indole dioxygenase (IDO1, IDO2) (EC 1.13.11.11 and 1.13.11.52), depending on the organism/tissue. 2. NFK is then converted to L-kynurenine (KYN) by the kynurenine formamidase (E.C. 3.5.1.9). 3. Finally, KYN is catalyzed to the unstable 4-(2-aminophenyl)-2,4-dioxobutanoate intermediate by a kynurenine transaminase (KAT1-4) (E. C. 2.6.1.7), which is readily converted to KYNA by a spontaneous reaction. The spontaneous cyclization of the intermediate to KYNA is unique to KYNA biosynthesis, and it makes this reaction effectively irreversible meaning that exogenous KYNA will not be converted to L-kynurenine nor its toxic derivatives.
Ijms 25 09082 g003
Figure 4. Transcript expression levels and Gini coefficient of SLC22A6 and SLC22A8. Data from [148,238].
Figure 4. Transcript expression levels and Gini coefficient of SLC22A6 and SLC22A8. Data from [148,238].
Ijms 25 09082 g004
Figure 5. 2D chemical structures of kynurenic acid and riboflavin, indicating a common substructure.
Figure 5. 2D chemical structures of kynurenic acid and riboflavin, indicating a common substructure.
Ijms 25 09082 g005
Figure 6. Assessing the fate of KYNA when its precursors are added externally is fraught unless one knows the expression levels of all the relevant transporters, the direction in which they transport, and whether they are concentrative or equilibrative [223,228], and we only know the existence of some of them. The membrane is redrawn in part from the animation at https://www.youtube.com/watch?v=s23vNwLE-Jw, accessed on 19 August 2024.
Figure 6. Assessing the fate of KYNA when its precursors are added externally is fraught unless one knows the expression levels of all the relevant transporters, the direction in which they transport, and whether they are concentrative or equilibrative [223,228], and we only know the existence of some of them. The membrane is redrawn in part from the animation at https://www.youtube.com/watch?v=s23vNwLE-Jw, accessed on 19 August 2024.
Ijms 25 09082 g006
Figure 7. A summary of the kynurenic acid concentrations of various foodstuffs. Data are compiled from the following references: squares [392], circles [36], diamonds [391].
Figure 7. A summary of the kynurenic acid concentrations of various foodstuffs. Data are compiled from the following references: squares [392], circles [36], diamonds [391].
Ijms 25 09082 g007
Figure 8. Production of reactive oxygen species as part of ischemia-reperfusion injury. Redrawn from the CC-BY 4.0 paper [413].
Figure 8. Production of reactive oxygen species as part of ischemia-reperfusion injury. Redrawn from the CC-BY 4.0 paper [413].
Ijms 25 09082 g008
Figure 9. Cheminformatic analysis of the structural similarity of kynurenic acid to those marketed drugs for which the Tanimoto similarity with the RDKit Pattern encoding exceeds 0.7.
Figure 9. Cheminformatic analysis of the structural similarity of kynurenic acid to those marketed drugs for which the Tanimoto similarity with the RDKit Pattern encoding exceeds 0.7.
Ijms 25 09082 g009
Figure 10. In silico prediction of the KYNA interactome. (A). Venn diagram analysis of predicted KYNA interacting proteins from the SuperPred [769], SwissTargetPrediction [768] and PharmMapper [767] cheminformatic tools. (B). Functional enrichment analysis of all (n = 455) predicted KYNA interactors. The DAVID [773] knowledgebase was used to identify the top functional annotation clusters, significantly enriched among the predicted KYNA interactome.
Figure 10. In silico prediction of the KYNA interactome. (A). Venn diagram analysis of predicted KYNA interacting proteins from the SuperPred [769], SwissTargetPrediction [768] and PharmMapper [767] cheminformatic tools. (B). Functional enrichment analysis of all (n = 455) predicted KYNA interactors. The DAVID [773] knowledgebase was used to identify the top functional annotation clusters, significantly enriched among the predicted KYNA interactome.
Ijms 25 09082 g010
Table 1. Some known substrates/inhibitors of ABCC4 (MRP4) that may assist in raising intracellular and tissue levels of KYNA.
Table 1. Some known substrates/inhibitors of ABCC4 (MRP4) that may assist in raising intracellular and tissue levels of KYNA.
MoleculeCommentsSelected References
Ceefourin-1Highly selective inhibitor of ABCC4[271,272,273]
Leukotrienes B4/C4Km 0.1–5.6 μM[274,275,276]
MK-571 (Verlukast)Inhibitor. Also a quinoline with a carboxylate group. Commonly more potent than probenecid. Also inhibits MRP1.[266,277,278,279]
ProbenecidInhibits multiple transporters.
Ki for SLC22A6/8~15 μM
[280,281,282]
QuercetinKi 1 μM (and other related polyphenols)[283]
Reviews [251,284,285,286]
SulindacKi 2 μM[287]
UratePertinent in gout and kidney disease[263]
Table 2. Some values for measured concentrations of KYNA in human body fluids, and rodent and human tissues.
Table 2. Some values for measured concentrations of KYNA in human body fluids, and rodent and human tissues.
Serum
Concentration RangeCommentsReferences
42 nM medianApproximately doubled in severe, acute COVID-19[331]
42 nM (35–54 nM IQR)Increase with age noted[332]
47 nM medianMarginally lower in ADHD[333]
38 nM meanUnchanged by inflammatory bowel diseases[334]
10-60 nM [335]
60 nM medianLower in ALS (ca 40 nM)[336]
23 nMNo different in pre-eclampsia[337]
38 nMHigher in erythrocytes of Parkinson’s disease, implying synthesis or concentration[338]
28 nM averageInsignificantly lower in depression[339]
Mean 25 nM, 16 nM in gestational diabetes (GD)L-kynureine was higher in the GD individuals[340]
100 nMMaternal serum; 3–4-fold higher in cord blood[341]
Plasma
Concentration rangeCommentsReferences
40 nM15% decrease in migraine[342]
~30 nM [343]
~40 nMNo effect in depression[344]
30–40 nMSmall increase with ibuprofen[345]
~40 nMIncreased 63% after endurance exercise (150 km road race)[215]
44 nM medianNo change in migraines[346]
10–80 nMMedian 38 nM; no difference in bipolar individuals[347]
39–54 nM medianSlightly greater with age and male gender[348]
~50 nMIncreased 3-fold when SLC22A6/8 inhibited by addition of probenecid[349]
18–350 nMPregnant women, 18–20 weeks, NB concentrated 20-fold in cord blood[294]
4–60 nMA summary of multiple measurements[350]
23 nM mean [351]
~20 nMChinese population[352]
70 nM in controls, 104 nM in those with social anxiety disorderIncreased with age in controls, but no relation with age in social anxiety disorder[353]
Mean 21 nM, 23 nM in women with pre-eclampsia (PE)Strongly influenced by BMI, that may have been a confounder; raised level suggested as a response to the PE rather than a cause[354]
CSF
Concentration rangeCommentsReferences
~20 nM in controlIncreased 3-fold in Alzheimer’s[355]
1–4 nMMedan 38 nM; no difference in bipolar individuals[336]
5 nMStable post mortem[356]
2–5 nMStrong positive correlation with age[357]
1–50 nMCan be raised strongly by certain alleles of KAT II[358]
Table 3. Some sources of kynurenic acid in certain foodstuffs.
Table 3. Some sources of kynurenic acid in certain foodstuffs.
VegetableCommentsSelected References
Broccoli0.41 mg/kg[307]
Chestnut honey129–601 mg/kg[392]
Chestnut honey~400 mg/kg[393]
Chestnut honeyUp to 2000 mg/kg[308]
Flower honeys (various)0.1–2 mg/kg[392]
Herbs of various kindsDandelion leaves 0.5 mg/kgww
St John’s wort 32 μg/dose
[386]
Horseshoe crab extract1200 mg/kg (0.12%)[360]
Potato tubers0.1–3.2 mg/kg[307]
0.3–3 mg/kg across 16 cultivars[391]
3× greater in purple potatoes[37]
Review[308]
Table 4. Some of the diseases or syndromes in which normal levels of KYNA are altered. Based in part on the data in [86], and see also the reviews [102,423].
Table 4. Some of the diseases or syndromes in which normal levels of KYNA are altered. Based in part on the data in [86], and see also the reviews [102,423].
Disease or SyndromeSource, and Raised or LoweredSelected References
Acute liver failureRaised in rat brain as a consequence, via increase in kynurenine in the periphery[424]
Alzheimer’s dementiaPlasma one third lower
Plasma marginal effect
[343,355]
Serum lower[425]
CSF lower[328,426]
CSF higher, but seemingly protective vs. disease progression[427]
Review[428]
Aortic stiffnessCorrelation of KYNA levels in patients with atrial fibrillation[429]
Attention deficit hyperactivity disorder (ADHD)Significantly lowered (meta-analysis of 650 individuals)[430]
Bipolar depression~40% decreased vs. controls[431]
Review; very variable, mostly lower[432,433]
CancersVery heterogeneous. Inhibition of proliferation observed at very high doses.[434,435,436]
Cluster headaches and migrainesSerum ~one third lower[437,438]
COVID-19Raised in serum, especially in more severe acute casesReviewed by [439,440], and see next section
Familial Mediterranean feverKYNA decreased[441]
FrailtyLowered in frailty
or little change
[329,442]
Huntington’s diseaseCerebral cortex—four-fold reduction.
One molecule (laquinimod) targeting the aryl hydrocarbon receptor in clinical trials
[443,444,445]
Inflammatory bowel diseaseSeen as a protective mechanism via raised levels of KATs[334,446]
Irritable bowel syndromeLowered in serum[447]
Lowered in urine[448]
Kidney diseaseNormal serum level of 28 nM increased to 336 nM in renal insufficiency[449]
Significantly raised in non-survivors of septic shock with acute kidney injury[450]
Chronic kidney disease, plasma levels can exceed 500 nM, along with high levels of other tryptophan metabolites[263,451]
End-stage kidney disease, plasma levels can exceed 1 μM[263]
Significantly raised in renal failure[452]
Major depressive disorder30% reduced over healthy controls; seen as the only metabolic biomarker that is both diagnostic and predictive.
~40% decreased over controls
Antidepressant activity in mice (at high concentrations)
Considered to be related to poor Western diet
[431,453,454,455]
Lowered in cortex, raised in serum of rats undergoing chronic restraint stress[456]
Migraine15% lower
KYNA/KYN halved
[342,457]
Significantly lower, acting via multiple receptors[458]
Multiple sclerosisPlasma 45 → 77 nM[459]
Erythrocytes 38 → 63 nM[459]
Raised in relapsing-remitting MS, lowered in primary and secondary progressive. Quinolinic acid seems to be the real culprit here.[460,461]
Lowered in CSF vs. other neurological diseases[462]
Myalgic encephalopathy/chronic fatigue syndrome (ME/CFS)Significantly lowered in some tissues, raised in others[463,464,465]
OsteoporosisPotential for treatment[466]
Parkinson’s disease (Review [467])Frontal cortex—~one third of controls[468]
Plasma[338]
Serum
Substantia nigra—less than half that of controls[468]
CSF lower[328]
Polycystic kidney diseaseSignificantly raised[469]
Polycystic ovary syndromeRoughly doubled[470]
Pre-eclampsiaVery nonlinear, but significantly raised (especially in those with high BMI) in one Norwegian birth cohort study[354]
No effect in a variety of other studies reviewed in:[337,471]
Pulmonary arterial hypertensionRaised, as were a great many other L-kynurenine pathway metabolites[472]
Schizophrenia and bipolar disorderRaised significantly (though usually so are other molecules such as L-kynurenine (which may be the real effector and/or changed by inhibitors of KATII)
Lowered in some studies
Meta-analysis implies no real or obvious difference
[196,198,199,433,473,474,475,476,477,478,479]
Sjögren’s syndromeKYNA somewhat raised[480]
Systemic lupus erythematosus (Lupus)Many kynurenine pathway metabolites raised[481]
Type 1 diabetesSteptozotocin-induced diabetes in rats led to a modest increase in KYNA[482]
Type 2 diabetesRaised from 36 to 46 nM[483]
Said to be raised during progression[484]
Ulcerative colitisRaised, seen as likely coming from changes in microbial gut metabolism.
Protection considered to be via GPR35
[377,485]
Table 5. Ability of KYNA to serve as a neuroprotective agent.
Table 5. Ability of KYNA to serve as a neuroprotective agent.
SystemCommentsSelected References
AnticonvulsantsPossible model of mediation via KYNA[509,510,511]
DepressionConsidered neuroprotective in depression[512]
Epileptic spasmsLower KYNA in epileptic spams that in other non-inflammatory neurological diseases[513]
Excitotoxic challengesKYNA protective[514,515]
Experimental autoimmune encephalomyelitisProtective against a Th17 response[379]
Ischemia-reperfusionGerbil brain. Massive doses led to very high intracerebral concentrations of KYNA and neuroprotection[516]
Highly protective in a model of hypoxic ischemia in neonatal rats[517,518]
Kynurenine sulphate produces KYNA that is neuroprotective in gerbils and rats[519,520]
Memory enhancementEffective at lower doses in mice
Opposite effect in C. elegans
[521,522]
MigraineSeems to be protective via inhibition of glutaminergic neurons[523,524,525,526]
Multiple sclerosisConsidered protective[527,528]
Plasma 45 → 77 nM[459]
Erythrocytes 38 → 63 nM[459]
PainProtective against neuropathic pain, and enhances effectiveness of morphine[529]
Antinociceptive in inflammatory pain[530]
Reviews [531,532,533]
Spinal cord injury (SCI)Protective (with glucosamine) against SCI in rats[534]
StrokeMostly protective if given ahead of experimental stroke.
Naturally protective against death after adjusting for inflammation
Higher levels associated with better recovery
[432,535,536]
Traumatic brain injuryKYNA is overexpressed, attenuates this in rats[537,538]
Table 7. Receptors to which KYNA has been suggested or found to bind.
Table 7. Receptors to which KYNA has been suggested or found to bind.
Putative ReceptorCommentsSelected References
Adrenoceptor alpha 2B (ADRA2B)—Putative ligandNote that guanfacine is an FDA-approved agonist, used successfully vs. attention deficit hyperactivity disorder (and interestingly also vs. hypertension [575])[576,577]
Review[578]
Identified in a high-throughput CRISPR screen[579]
Aryl hydrocarbon receptor (AhR)—AgonistInduces various pathways, including IL-6 production, at 100 nM[580,581]
Protects against intestinal C. albicans infection via AhR[582]
Possible role in COVID-19[439,583]
Protective against acute lung injury[562]
Removing AhR raises KYNA levels, and these are neuroprotective against excitotoxic insults[93]
KYNA affect neural plasticity via AhR in zebrafish[584]
Involved in fibrosis and
skin disease
[585]
[586]
G-protein-coupled receptor 35 (GPR35)—AgonistRegulates energy metabolism and ablates weight gain on high-fat diet in mice[92,213,360,587]
Protection against ischemic injury (at 5 mg/kg, ≡26 μM if homogeneous)[588]
Reviews[589,590,591]
Glutamate receptor (GAR)—antagonist Many papers relating to migraines, e.g.,[523,524,525,526]
Hydroxycarboxylic acid receptor 3 (HCAR3)—putative ligandIdentified in a high-throughput CRISPR screen[579]
N-methyl D-aspartate (NMDA) receptor (especially the strychnine-sensitive glycine-binding site) (involved in pain [592])—antagonistEC50 7 μM
though binding curves and very complex effects
[69,593,594]
Potential role in nutritional signalling[595]
10 nM can affect differentiation of cortical cells[596]
Electrophysiological effects observable at high concentrations[597]
Reviews[598,599,600]
alpha-7 nicotinic acetylcholine receptor (α7nAChR)—purported antagonistElectrophysiological effects not observed even at high concentrations[597]
Active at 7 μM in hippocampal neurons[601]
No physiological effects observed with KYNA[602]
Lowers inflammatory cytokine production and Abeta phagocytosis[603]
Table 8. Some circumstances in which exogenous elements affect the levels of KYNA.
Table 8. Some circumstances in which exogenous elements affect the levels of KYNA.
SubstanceCommentsSelected References
AmphetaminesSignificant decrease after dosing with amphetamine[616]
Anthocyanins (in blackberry extract)Microbiome said to be responsible for increasing KYNA levels (though in the LC-MS data neither the apparent retention time nor the reported mass of the positive molecular ion (m/z = 208; true is 190) underlying this is that of KYNA)[617]
Antidiabetic agentsGlibenclamide and metformin both decrease KYNA levels, likely by different mechanisms[618]
ExerciseExercise can stimulate the production of katG enzymes and thereby raise KYNA (and lower central L-kynureine), making KYNA an ‘exerkine’. Data are somewhat mixed[619,620,621]
KAT4 was especially strongly stimulated in endurance exercise, leading in some cases to more than a 60% increase in plasma KYNA levels[215]
Exercise increases KYNA and its activation of the AhR receptor[622]
Fasting8 day fasting increased KYNA levels.
Effect observable even at 2 days
[623,624]
HypothyroidismExperimentally induced hypothyroidism leads to brain KYNA levels being raised[625]
Insulin signallingTested in C. elegans, caused increases in KYNA[522]
Interferon-γMassive increase in KYNA in neurons and astrocytes[626,627]
Ketone bodiesβ-hydroxybutyrate stimulates KYNA synthesis and provides neuroprotection[514,515,628,629]
LPS treatmentLowers brain KYNA[630]
ObesityHigher serum levels of several KP metabolites, including KYNA; in some cases, may simply reflect higher fluxes from raised dietary intake[631]
Estrogens (as oral contraceptive agents)Decreased from a median of 58 → 33 nM in plasma[632]
Progesterone partial reverses interferon-γ-induced decrease in KYNA[633]
Hericium erinaceus polysaccharidesHepatoprotective vs. non-alcoholic fatty liver disease, and increase KYNA levels[634]
StressStress increases KYNA levels in rats, who show lower cognitive skills; unclear whether KYNA response is causally involved or an attempt to counteract. KYNA was not added independently
Hepatic KYNA lowers anxiety-induced stress
[635,636]
Table 9. Some other known biochemical and physiological effects of KYNA.
Table 9. Some other known biochemical and physiological effects of KYNA.
PathwayCommentsSelected References
Amyloid (Aβ) fibrillation and toxicity in C. elegansInhibited by KYNA, as it was by some simple analogues[637]
Anti-inflammatoryIncludes effects on histone methylation[638]
Reverses effects of LPS in macrophage cultures[639]
Lowers inflammatory phagocytic response in mouse macrophages[640,641]
Lowers LPS-induced inflammation[642]
Lowers experimentally induced inflammation in the trigeminal ganglion[643]
Hydrogels containing KYNA lowers experimentally induced inflammation[644]
Anxiolytic (reduces anxiety)Notable effects in zebrafish at 105 μM (20 mg/L)[645]
ApoptosisInduced by KYNA[646]
270 genes differentially expressed after exposure to 0.25 mM KYNA[647]
Astrocyte activationInhibitory and protects against HIV-induced cognitive loss[648]
DNA excision repairKYNA increases pathway transcription[649]
Endothelial damage (induced by homocysteine)Protective[650]
Enhances endothelial adhesion and spreading[651]
Fibroblast growth factor release from HUVEC cellsInhibited at “low” concentrations (1 μM) of KYNA, while proliferation rate increased.[652]
Glutamate releaseLowered by KYNA[653]
HypertensionHeart rate lowered by KYNA in spontaneously hypertensive rats[654]
Indoleamine 2,3-dioxygenase inductionSignalling role[562,655]
Insulin resistanceProtective in high-fat-diet-induction model[656]
InterleukinsLowered IL17/IL23 at high concentrations[657]
Mitochondrial inductionActs as a cardioprotective[557]
NeprilysinNeprilysin degrades amyloids, and KYNA induces its synthesis and is neuroprotective[658]
TNF-α productionDecreased at very high concentrations[659,660,661]
Unfolded protein responseKYNA inhibits, and is protective in a C. elegans Alzheimer’s model[662]
VasculatureInduces vascular relaxation in endothelial cells[663]
Table 10. Some studies in which organisms or mammalian cells have been exposed to exceptionally high concentrations of KYNA.
Table 10. Some studies in which organisms or mammalian cells have been exposed to exceptionally high concentrations of KYNA.
OrganismDose and CommentSelected References
Gerbils400–1600 mg/kg; protected against ischemia-reperfusion injury[516]
Mice250 mg/L in drinking water, ≡25 mg/kg/d, has no toxic effects[640,641]
25–250 mg/L drinking water 3–21 d; well tolerated.[665]
Rat500 mg/kg i.p.—protected against thioacetamide-induced liver injury[666]
300 mg/kg i.p. in young rats prolonged wakefulness[323,667]
150 mg/kg lowers morphine-conditioned reward behavior[668]
25 mg/kg/d in drinking water ≡ ~250 mg/L, assists healthy growth in rat babies[359]
300 mg/kg i.p. protect against mussel toxin[554]
200 mg/kg kynuramine[669]
300 mg/kg vs. acute pancreatitis[564]
300 mg/kg in young rats was neuroprotective[517,520]
As much as 5% of diets, with large KYNA excretion, small decrease in weight gain, but seemingly without major ill effects.[670,671]
Mammalian cell lines
Murine RAW 264.7 macrophages100 μM KYNA is protective against LPS challenges[639]
Murine BV-2 microglial cellsNo effect on viability of 100 μM KYNA[603]
Rat splenocytesNo effects on viability of proliferation at 500 μM[641]
HumansTopical application; no safety issues observed. Most secreted in urine[551]
Table 11. Some examples in which KYNA was considered to have some potentially negative effects.
Table 11. Some examples in which KYNA was considered to have some potentially negative effects.
General Biological AreaCommentsSelected References
Cardiovascular diseaseMarginally associated with (a 10% increase) in all-cause, but likely confounded with L-kynurenine that is more significant[568]
SchizophreniaKYNA levels often raised, though unclear if this is a cause or an attempted detoxification of raised L-kynurenine. Some studies showed them lowered; overall unclear. See also the text.[196,473,474,476,478,684]
Table 12. KYNA protein targets predicted by two or more computational cheminformatics tools.
Table 12. KYNA protein targets predicted by two or more computational cheminformatics tools.
Protein NamesUniProt IDEntry NamePredicted to Bind KYNA by
Thyroid hormone receptor alpha P10827THA_HUMANSuperPred, SwissTarget, PharmMapper
Glutathione S-transferase P P09211GSTP1_HUMANSuperPred, PharmMapper
Inosine-5′-monophosphate dehydrogenase 2 P12268IMDH2_HUMANSuperPred, PharmMapper
Galectin-3 P17931LEG3_HUMANSuperPred, PharmMapper
Histone deacetylase 8 Q9BY41HDAC8_HUMANSuperPred, PharmMapper
Gamma-aminobutyric acid receptor subunit alpha-1 P14867GBRA1_HUMANSuperPred, SwissTarget
D-amino-acid oxidase P14920OXDA_HUMANSuperPred, SwissTarget
Amine oxidase (flavin-containing) A P21397AOFA_HUMANSuperPred, SwissTarget
Aryl hydrocarbon receptor P35869AHR_HUMANSuperPred, SwissTarget
Dual specificity protein kinase CLK4Q9HAZ1CLK4_HUMANSuperPred, SwissTarget
Prothrombin Thrombin heavy chainP00734THRB_HUMANSwissTarget, PharmMapper
Renin P00797RENI_HUMANSwissTarget, PharmMapper
Carbonic anhydrase 1P00915CAH1_HUMANSwissTarget, PharmMapper
Carbonic anhydrase 2 P00918CAH2_HUMANSwissTarget, PharmMapper
Thymidylate synthaseP04818TYSY_HUMANSwissTarget, PharmMapper
Lymphocyte specific tyrosine kinase P06239LCK_HUMANSwissTarget, PharmMapper
NeprilysinP08473NEP_HUMANSwissTarget, PharmMapper
Leukotriene A-4 hydrolase P09960LKHA4_HUMANSwissTarget, PharmMapper
Thyroid hormone receptor beta P10828THB_HUMANSwissTarget, PharmMapper
Angiotensin-converting enzymeP12821ACE_HUMANSwissTarget, PharmMapper
Farnesyl pyrophosphate synthaseP14324FPPS_HUMANSwissTarget, PharmMapper
Neutrophil collagenase P22894MMP8_HUMANSwissTarget, PharmMapper
Macrophage metalloelastase P39900MMP12_HUMANSwissTarget, PharmMapper
Aldo-keto reductase family 1 member C3P42330AK1C3_HUMANSwissTarget, PharmMapper
Mitogen-activated protein kinase 10 P53779MK10_HUMANSwissTarget, PharmMapper
Peroxisome proliferator-activated receptor alphaQ07869PPARA_HUMANSwissTarget, PharmMapper
Estrogen receptor beta (ER-beta) Q92731ESR2_HUMANSwissTarget, PharmMapper
Table 13. Fermentative production of KYNA in microorganisms.
Table 13. Fermentative production of KYNA in microorganisms.
OrganismGenetic Modification(s)TiterConditions and CommentsReferences
Escherichia coliRemove competing pathway, enhance SAM pathway350 mg/mLMain target was actinocin[780]
Saccharomyces cerevisiaeNone, but a chemical defined medium including 400 mg/L tryptophan was used9 mg/L [781]
~1.5 μM (~280 ng/mL) [782]
Yarrowia lipolyticaNone21 μg/mL in culture broth or 494 μg/g cell dry weight trp-supplemented media[783,784,785]
Table 14. A comparison of some properties ergothioneine and KYNA in terms of our knowledge and their present status as nutraceuticals.
Table 14. A comparison of some properties ergothioneine and KYNA in terms of our knowledge and their present status as nutraceuticals.
PropertyErgothioneineKynurenic Acid
Overall status as a nutraceuticalFairly well established [1,18,29,508,791,792,794,795]Emerging [86,99,308,533]
Biosynthesized endogenouslyNoYes
Largest dietary sourceMushrooms (fruiting bodies of basidiomycetes) [29,796,797,798,799]Chestnut honey [308,392]
Highest natural product level ~9 g/kg [29,800]~0.6 g/kg [308,392]
Approximate serum/plasma level in ‘healthy’ populations~260 ng/mL, ~1.1 μM ~40 nM
Degree of concentration in erythrocytesMaybe as much as 10 times [801]Possibly 3× but unclear how calculated [425]; no increase in a related study [459]
Known ‘uptake’ transporters in humansRelatively specific and concentrative e.g., SLC22A4 [802,803], SLC22A15 [804]Those known are non-specific and non-concentrative [235,236]
Concentrated in erythrocytes as a kind of ‘buffer’ for plasmaSignificantly (up to ten-fold [801,805]), probably accumulated via SLC22A4 [806]Much less so (note that RBC express ABCC4 [807])
Known ‘efflux’ (pump) transporters in humansNot knownABCC4 [242,696] and possibly ABCG2 [263,264]
Known transporters in microorganismsSeveral, e.g., [778,808,809]None seemingly published
ThermostabilityHigh (can be extracted at 95 °C [810])High (stable to boiling and frying [37]), and in blood samples [699]
Pharmacokinetics studiesSomewhat, with human feeding trials for 35 d [319]Not really initiated orally
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

Alves, L.d.F.; Moore, J.B.; Kell, D.B. The Biology and Biochemistry of Kynurenic Acid, a Potential Nutraceutical with Multiple Biological Effects. Int. J. Mol. Sci. 2024, 25, 9082. https://doi.org/10.3390/ijms25169082

AMA Style

Alves LdF, Moore JB, Kell DB. The Biology and Biochemistry of Kynurenic Acid, a Potential Nutraceutical with Multiple Biological Effects. International Journal of Molecular Sciences. 2024; 25(16):9082. https://doi.org/10.3390/ijms25169082

Chicago/Turabian Style

Alves, Luana de Fátima, J. Bernadette Moore, and Douglas B. Kell. 2024. "The Biology and Biochemistry of Kynurenic Acid, a Potential Nutraceutical with Multiple Biological Effects" International Journal of Molecular Sciences 25, no. 16: 9082. https://doi.org/10.3390/ijms25169082

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