Next Article in Journal
Insulin-like Growth Factor 1, Growth Hormone, and Anti-Müllerian Hormone Receptors Are Differentially Expressed during GnRH Neuron Development
Previous Article in Journal
Defining Mechanisms of C3 to CAM Photosynthesis Transition toward Enhancing Crop Stress Resilience
Previous Article in Special Issue
Critical Roles of the Cysteine–Glutathione Axis in the Production of γ-Glutamyl Peptides in the Nervous System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Glutathione-Mediated Neuroprotective Effect of Purine Derivatives

Department of Pharmacology, Teikyo University School of Medicine, 2-11-1 Kaga, Itabashi, Tokyo 173-8605, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(17), 13067; https://doi.org/10.3390/ijms241713067
Submission received: 16 June 2023 / Revised: 19 August 2023 / Accepted: 20 August 2023 / Published: 22 August 2023
(This article belongs to the Special Issue Neuroprotective Effect of Glutathione 2.0)

Abstract

:
Numerous basic studies have reported on the neuroprotective properties of several purine derivatives such as caffeine and uric acid (UA). Epidemiological studies have also shown the inverse association of appropriate caffeine intake or serum urate levels with neurodegenerative diseases such as Alzheimer disease (AD) and Parkinson’s disease (PD). The well-established neuroprotective mechanisms of caffeine and UA involve adenosine A2A receptor antagonism and antioxidant activity, respectively. Our recent study found that another purine derivative, paraxanthine, has neuroprotective effects similar to those of caffeine and UA. These purine derivatives can promote neuronal cysteine uptake through excitatory amino acid carrier protein 1 (EAAC1) to increase neuronal glutathione (GSH) levels in the brain. This review summarizes the GSH-mediated neuroprotective effects of purine derivatives. Considering the fact that GSH depletion is a manifestation in the brains of AD and PD patients, administration of purine derivatives may be a new therapeutic approach to prevent or delay the onset of these neurodegenerative diseases.

1. Introduction

A variety of purine derivatives are produced from purine nucleotides (i.e., adenine nucleotide and guanine nucleotide) by metabolic processes occurring mainly in the liver. For instance, adenine nucleotides are metabolized to produce ATP, ADP, AMP, adenosine, inosine, hypoxanthine, and xanthine, and finally uric acid (UA). Caffeine (1,3,7-trimethylxanthine), a naturally occurring purine derivative, is metabolized into several dimethylxanthines including paraxanthine, theophylline, and theobromine, and finally into UA derivatives.
Each purine derivative has a particular physiological activity, and several purine derivatives such as adenosine, guanosine, caffeine, paraxanthine, theophylline, theobromine, and UA have been shown to possess neuroprotective activities [1,2,3,4,5,6,7,8,9]. Caffeine has an adenosine A2A receptor (A2AAR) antagonizing activity that is involved in neuroprotection [5]. The neuroprotective activity of UA is attributed to its antioxidative activity [10,11]. The neuroprotective activities of both caffeine and UA are supported by several epidemiological studies, which show negative correlations of dietary caffeine intake and serum UA levels with the onset of neurodegenerative diseases such as Alzheimer disease (AD) [12,13,14,15] and Parkinson’s disease (PD) [16,17,18,19]. Recently, we have shown that caffeine and its metabolites, paraxanthine and UA, have neuroprotective activities, enhance cysteine (Cys) uptake, and increase intracellular glutathione (GSH) levels [20,21]. GSH is an intracellular antioxidant tripeptide molecule (glutamylcysteinylglycine), which plays an important role in neuronal survival under oxidative stress in the central nervous system (CNS) [22,23]. Therefore, it is likely that the increase in GSH levels is related to the neuroprotective activities of purines such as caffeine, paraxanthine, and UA [20,21].
Increasing GSH levels is a neuroprotective mechanism that increases antioxidant activity in neurons. It is widely accepted that neuroprotective processes can be classified into three major mechanisms: (i) suppressing excessive neuronal stimulation, (ii) maintaining antioxidant activity in neurons, and (iii) detoxification of xenobiotics (Table 1). Intracellular GSH is directly or indirectly involved in all four of the intracellular antioxidizing pathways, i.e., Table 1 (ii, 1–4): intracellular GSH synthesis, Cys uptake, GSH supply from astrocytes, and antioxidant supply.
Figure 1 summarizes the regulation of GSH synthesis in neurons and astrocytes. GSH is synthesized from glutamate (Glu), Cys, and glycine (Gly) by two-step enzymatic reactions [32,49]. The first step of GSH synthesis is catalyzed by γ-glutamyl cysteine ligase (GCL), which is composed of a GCL catalytic subunit (GCLC, Gene ID: 2729) and a GCL modifier subunit (GCLM, Gene ID: 2730) [30,31,32]. The second step of GSH synthesis is catalyzed by GSH synthetase (GSS, Gene ID: 2937) [33]. The GSH synthesis process is preceded by Cys uptake, which is mediated by excitatory amino acid carrier protein 1 (EAAC1, Gene ID: 6505) in neurons [35,36]. Cys uptake into neurons is a major pathway that supplies material for GSH production. Transporting Cys is mainly mediated by a neuron-specific Cys transporter, EAAC1 [34,35,50].
The intracellular presence of the antioxidants such as ascorbate, UA, and α-tocopherol help to preserve neuronal GSH levels by doing the work of protecting cells against oxidative stress through other means [11,40,41,42,43]. Neuronal GSH levels are also affected by GSH production in the astrocytes that surround and support the neurons [28,37,51]. Astrocytes release GSH via multidrug resistance protein 1 (MRP1, Gene ID: 4363); the GSH is degraded extracellularly and then taken up by neurons to be reconstructed [39,52]. This GSH release is promoted by enhancing GSH synthesis in astrocytes, which is mainly regulated by GCL, GSS, and cystine (cysteine disulfide) uptake via the cystine/glutamate antiporter, system xc (xCT, Gene ID: 23657) [38] (Figure 1).
Several purine derivatives are found to be neuroprotective, and they increase intracellular GSH levels not only in neurons but also in astrocytes. Thus, the regulation of GSH levels is likely integral to the neuroprotective activity of purines in the CNS. We present here the relation of purine derivatives with neuroprotection, especially in terms of GSH synthesis (in Section 3.3).
Whereas increased GSH is known to be neuroprotective, reduced brain GSH levels have been reported to precede the pathologic hallmarks of AD such as amyloid oligomerization and plaque formation in AD model mice [53]. In a clinical study, GSH depletion was considered an early event in the progression of PD [54]. Thus, promoting intracellular GSH synthesis prior to symptomatology may halt the progression of neurodegenerative diseases such as AD and PD. Basic research on the effect of purine derivatives on increasing GSH levels is valuable for developing novel disease-modifying drugs for the treatment of neurodegenerative diseases.
In this review, we summarize the role of purine derivatives in enhancing neuroprotective activities and alleviating neurodegenerative insults. In the second section, we describe recent epidemiological studies on the relationship of both caffeine and UA to the onset of neurodegenerative diseases. Then, in the final section, we present the neuroprotective functions of purine derivatives especially in terms of GSH synthesis.

2. Epidemiological Studies of the Relation between Caffeine or UA and Lower Risks of Neurodegenerative Diseases

Aging societies such as that in Japan face serious problems of age-related neurodegenerative disease such as AD and PD and especially the cognitive decline that accompanies these diseases. No effective curative or prophylactic treatment for the development of AD or PD has so far been clinically developed. Genetic backgrounds are involved in the onset and progression of familial AD and PD [55,56], while several environmental factors that promote or attenuate the onset of sporadic AD [57] and PD [58,59] have been identified in epidemiological studies. Previous studies have suggested the involvement of factors such as caffeine intake and serum UA levels in modulating the incidence or progression of AD and PD. Serum caffeine concentrations in PD patients and matched healthy control are about 2.4 µM and 7.9 µM, respectively [60]. UA concentration in PD patients and matched healthy control is about 274 µM and 286 µM, respectively [19]. A higher serum UA (≥271 µM) is associated with lower risk for AD compared to lower serum UA (≤210 µM) [15]. The neuroprotective activities of both caffeine and UA have been confirmed by preclinical studies using animal models. These studies suggest that the neuroprotective mechanisms of caffeine and UA are respectively involved in A2AAR antagonism (Kb = 12.3 µM, Table 2) [3,61,62] and in antioxidant activity (200 to 500 µM UA) [63,64], namely, with upregulation of the signaling pathway that produces antioxidative molecules in cells.

2.1. Relation between Caffeine Intake and Lower Risk of Neurodegenerative Diseases

Since the 1980s, numerous epidemiological studies have shown a correlation between coffee/caffeine intake and lowering the risk of neurodegenerative diseases such as AD and PD. A negative correlation was found between caffeine intake and the incidence or progression of AD [12,13,16]. In one study, caffeine intake was correlated with a lower risk of cognitive decline in women, but not significantly so in men [12]. To clarify whether the coffee-induced effect could be attributed to molecules other than caffeine, Dong et al. [13] reported the contribution of caffeine to reducing the risk of cognitive decline in elder adults by comparing five different groups, including those who did not consume coffee, and those who consumed coffee, caffeinated coffee, decaffeinated coffee, and caffeine from coffee. The results showed that cognitive performance was correlated with the intake of coffee, caffeinated coffee, and caffeine from coffee, but not decaffeinated coffee. Thus, these findings suggest that intake of caffeine, rather than other components of coffee, significantly lowers the risk of cognitive decline of AD patients.
Coffee/caffeine intake has also shown correlations with better cognitive performance in PD patients [16,17,19,70,71,72,73,74,75,76,77,78,79]. Although the correlation between coffee intake and the lower risk of PD was not significant in early studies [70,71], subsequent larger case-control studies demonstrated the relation of a lower risk of PD to coffee/caffeine intake [16,17,19,72,73,74,75,76,77,78,79]. Ross et al. [17] showed that coffee/caffeine intake was correlated with a low incidence of PD, whereas tobacco use (smoking) and administration of other nutrients contained in coffee did not lower the risk of PD.
More than 90% of caffeine clearance is mediated by cytochrome P450 family 1 subfamily A member 2 (CYP1A2, Gene ID: 1544) [80]. A polymorphic variant of CYP1A2 (–163 C > A) (GenBank accession number AF253322) confers higher CYP1A2 inducibility and higher individual caffeine metabolic activity [81]. Recently, Tan et al. [82] examined the correlation of caffeine intake with PD risk in terms of caffeine metabolism: in their case-control study, there was no difference in the relationship of caffeine intake to low risk of PD between fast and slow caffeine metabolizer genotypes. Even after normalizing caffeine absorption and metabolism, reduced salivary caffeine levels in PD patients correlates with PD progression [83]. These results further support the neuroprotective activities of both caffeine and its major metabolite, paraxanthine, found in animal studies [6,21,84]. However, it remains unclear whether paraxanthine alone is correlated with a lower risk of PD.

2.2. Relation between Serum UA and Risk of Neurodegenerative Diseases

High serum UA levels cause some critical diseases such as gout, cardiovascular disease, hypertension, and renal disease [85], while it has neuroprotective effects in PD and AD animal models, because UA has antioxidant activity. Several clinical studies have demonstrated the neuroprotective effect of high serum UA levels on neurodegenerative diseases such as PD and AD [19,86].
Although some conflicting results have been reported [87], a significant correlation between serum UA levels and a low risk of PD have been shown by many studies [19,88,89,90,91,92]. Some of the studies identify a difference in the UA/PD relationship between men and women [18,93]. In one study, higher plasma UA levels correlated with lower risk of PD in men but not in women [93]. Cortese et al. [18] also demonstrated that higher serum UA levels are related to lower risk of PD in men. Although the relation is weaker in women, the protective effect of serum UA is significantly increased in aged women (above 70 years), whose UA levels are higher than those in premenopausal women. These studies further support the correlation between serum UA levels and lower risk of PD.
A strong correlation of plasma antioxidant levels with both mild cognitive impairment and AD has been observed [14,86]. Furthermore, it has been shown that serum UA levels are correlated with lower risks of dementia, AD, and vascular dementia [15]. Interestingly, in participants without dementia, there was a correlation between higher serum UA levels and better cognitive performance in life; however, in participants with dementia, there was a correlation of high serum UA levels with declines in cognitive performance and manifestation of brain atrophy [94].
As mentioned above, some studies have supported the potential role of hyperuricemia to prevent AD, while other studies have reported conflicting results or have failed to demonstrate any significant association [95,96]. Although the results of these clinical studies have been inconsistent, a recent meta-analysis supported the correlation between serum UA levels and scores on the Mini-Mental State Examination in patients with PD-related dementia [14,15,94,97].
By treatment with urate-lowering agents, patients with gout maintain lower serum UA levels over long periods of time. One might think that low serum UA levels would impair neurons by increasing oxidative stress in the brain. However, use of urate-lowering drugs (allopurinol, febuxostat) is not correlated with any increase in the risk of dementia [95]. Treatment with pegloticase, a PEGylated urate oxidase (uricase), reduced mean serum UA levels more than 90% (from 10.8 to 0.9 mg/dL = 642 to 53 μM); however, there was no significant increase in biomarkers of oxidative stress, and the levels of oxidative markers did not correlate with serum UA levels [98]. Thus, drug-controlled low serum UA levels do not appear to affect the development of dementia or oxidative stress. The lack of correlation between low levels of serum UA and the development of dementia may be due to the absence of elevated oxidative stress levels. Indeed, it is not yet fully understood whether UA acts as an antioxidant in blood [11,99] and how serum UA provides neuroprotection in the CNS.

3. Neuroprotective Activities of Purines

3.1. Regulations of UA Levels in Blood and the Brain

Serum UA levels are maintained by food digestion, purine synthesis, metabolism, and purine excretion into urine. The net excretion of UA in urine is determined by the balance between UA re-absorption and secretion within the proximal tubule, and each process is mediated by its specific transporter (Figure 2). Since uricase activity is absent in primates, UA is the end metabolite of purines in humans. Therefore, serum UA levels in humans are higher than in other mammals. Abnormally high UA levels are related to the onset of diseases such as gout, cardiovascular disease, hypertension, and renal disease [85] because UA behaves as a pro-oxidant under certain circumstances: for example, in the presence of transition metals in the microenvironment [100]. However, higher UA levels in serum and cerebrospinal fluid have been effective in protecting neurons from oxidative stress in animal studies [8,9,101] and have been related to a lower onset of AD and PD in clinical studies (as described in Section 2).
UA is also produced during ATP metabolism through the following steps [102]: (1) ATP is converted to ADP and AMP by nucleoside triphosphate diphospho-hydrolases (NTPDases; CD39, Gene ID: 953) [103], (2) AMP is converted to adenosine by 5′-nucleotidase (5′NT; CD73, Gene ID: 4907) [104], (3) adenosine is converted to inosine by adenosine deaminase (ADA, Gene ID: 100), (4) inosine is converted to hypoxanthine by purine nucleoside phosphorylase (PNP, Gene ID: 4860), and (5) xanthine oxidase (XO, Gene ID: 7498) catalyzes the conversion of hypoxanthine to xanthine and finally to UA [105,106] (Figure 2). In mammals, XO is abundant in the liver, intestine, and mammary gland whereas it is scarce in the heart, muscle, and brain [105]. In immunolocalization studies, XO protein is located in mammary epithelial cells and in the capillary endothelial cells of almost all tissues except the brain and testis in bovines. The human brain lacks XO, and the XO activity is absent from neurons, astrocytes, epithelial cells, endothelial cells, and capillary endothelial cells in the brain [105]. Importantly, because UA production mainly occurs in the liver and not in the brain, serum UA can be supplied to the brain only by passing through the blood–brain barrier (BBB) (Figure 3).
In 1981, Granger et al. and McCord et al. [107,108] hypothesized that XO-generated reactive oxygen species (ROS) cause ischemic reperfusion injury in bowel and cardiac tissue. AMP catabolism to hypoxanthine occurs under hypoxic conditions in the ischemic process, while in the reperfusion process, XO and O2 mediate hypoxanthine oxidation to form xanthine and the end metabolite, UA, concomitant with super oxide anions (O2). Therefore, the lack of XO in the brain may protect neurons from such oxidative stress. In vivo preclinical study has shown that UA protects hippocampal neurons after ischemia reperfusion in rats [109].
In terms of the mechanism by which UA in blood is involved in neuroprotection in the brain, Amaro et al. [110] argued that the main target of UA is endothelial cells, because UA is practically unable to pass through the BBB. The protection of endothelial cells from ischemic injury leads to the survival of the whole neurovascular unit. When the BBB is impaired, UA passes through it. In fact, UA levels in cerebrospinal fluid correlate positively with serum UA levels, especially when the BBB is impaired [96]. The neuroprotective activity of UA has been reported in 6-hydroxydopamine (6-OHDA) lesioned PD model rats [8], as in ischemia-injured rats [9,109]. Although it is not clear how UA crosses the BBB and how it acts on the neurons, these results support the notion that systemic administration of UA might induce neuroprotective activity in the CNS.

3.2. Regulation of Purine Metabolism in the Brain

Other than nucleic acids, the typical intracellular purine derivative is ATP, which acts as an energy source for driving neuronal activity. Huge amounts of ATP (on the order of several mM) are produced in nervous tissues to maintain energy for Na+/K+ ATPase and synaptic transmission [111,112,113]. In contrast, extracellular ATP (at a concentration of low nM) regulates cellular signaling via P2-purinergic receptors such as P2Y and P2X [114,115,116] (Figure 3). ATP is released from cells under physiological conditions [117] and is also released from damaged cells [118,119,120] (Figure 3).
ATP is converted to ADP, AMP, adenosine, inosine, hypoxanthine, and xanthine and finally yields UA in the peripheral tissues (Figure 2). These purine derivatives have a variety of physiological activities by modulating adenosine receptors in the brain. Methylxanthines (MX) such as caffeine, paraxanthine, theophylline, and theobromine can be converted to UA by both demethylation and oxidation processes (Figure 2).
Adenosine is formed from AMP intracellularly in the CNS. Under physiological conditions, ATP and adenosine can be released from presynaptic neurons. Adenosine is also produced in extracellular space from AMP derived from released ATP and cyclic AMP (cAMP) [121,122,123]. In addition, astrocytes in the brain use ATP to regulate the extracellular concentration of purines, the metabolism of which is catalyzed by ecto-enzymes on cell membranes [122,124]. ATP is rapidly converted to ADP, AMP, adenosine, and inosine in the extracellular space [117,120,125] (Figure 3). ATP is decomposed into ADP and AMP by CD39 [103], AMP is decomposed into adenosine by CD73 [126], and adenosine is decomposed into inosine by ADA [127]. Adenosine stimulates adenosine A1 receptor (A1AR) and A2AAR (EC50 values for A1AR and A2AAR are 0.31 µM and 0.73 µM, respectively, Table 2). Trauma and ischemia induce the expression of CD39 and CD73 in astrocytes and increase extracellular adenosine derived from ATP metabolism [126,128] (Figure 3). Clearly, a variety of purine derivatives are produced from metabolic processes of purine nucleotides, and at least some of them seem to have neuroprotective activity in vitro and in vivo.

3.3. Neuroprotective Mechanisms of Purines

As described in the Introduction and illustrated in Table 1, neuroprotective processes can be classified into three major categories (Table 1): (i) suppressing excessive neuronal stimulation by neurotransmitter antagonism, (ii) maintaining antioxidant activity in neurons, which limits cellular oxidative stress caused by electrophile molecules such as ROS, and (iii) detoxifying xenobiotics such as nucleophile molecules and abnormally aggregated proteins such as α-synuclein and amyloid β peptide. Known neuroprotective activities induced by purine derivatives have been classified in categories (i) and (ii) above, but category (iii)-mediated neuroprotective effects of purine derivatives have been unclear.

3.3.1. Neuroprotection by Antioxidative Activity

UA has antioxidant activity [11,129,130] that leads to an established neuroprotective property, while xanthine structure has no such activity [64]. Structure-activity relationships for purine derivatives indicate that the 8-one substituent in the chemical structure of UA plays an important role in its antioxidant properties [63,129]. It has been suggested that UA may neutralize reactive oxygen species produced via a Fenton-type chemical reaction in cells. However, UA levels in cerebrospinal fluid (17.7 µM) are lower than those in plasma (172.3 µM) [96] (Figure 3), suggesting that UA might have an antioxidant effect in blood.
UA both acts as a scavenger of ROS and peroxynitrite [99] and prevents iron-mediated ascorbate oxidation [131]. In vitro, UA has an antioxidant activity similar to that of ascorbate, but humans have much higher levels of UA than ascorbate, due to their loss of uricase function during the course of evolution. However, low serum UA levels do not increase oxidative stress in blood [98]. In fact, UA is a less effective antioxidant than ascorbate in human blood [10,11]. This is because of UA’s alternative role as an iron chelator. Therefore, UA inhibits iron-catalyzed oxidation of ascorbate and stabilizes ascorbate levels in serum [132]. Antioxidant activity is observed at UA concentrations of more than 200 µM [10,64]; however, a much lower concentration of UA (10 µM) was shown to increase Cys uptake through the Cys transporter, resulting in GSH synthesis in hippocampal slices [20]. Thus, UA may contribute neuroprotection not only by antioxidant activity but also by promoting GSH production in neurons.
The neuroprotective activities of UA and caffeine have been confirmed by several preclinical studies. The well-established neuroprotective mechanism of UA is its antioxidant activity (10 to 100 µM), while that of caffeine is A2AAR receptor antagonism (IC50 = 107 µM, Table 2). Furthermore, UA and caffeine have other neuroprotective mechanisms in which GSH levels are increased by upregulating expressions of antioxidant-related proteins (Figure 3 and Table 3). UA appears to increase GSH levels by enhancing expression of nuclear factor erythroid-2-related factor 2 (Nrf2)-responsive genes, including GCLC, heme oxygenase-1 (HO-1), and NAD(P)H quinone oxidoreductase 1 (NQO1), and the resultant GSH increase provides antioxidant activity (Table 3).
There are numerous examples of UA’s direct neuroprotective roles. UA protects PC12 cells from 6-OHDA-induced cell injury by increasing levels of both GSH and superoxide dismutase (SOD) protein [133]. UA treatment enhances SOD activity, increases GSH levels, and reduces oxidative products of malondialdehyde (MDA) in a 6-OHDA-induced PD model rat [8]. In 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-treated PD model mice, UA improves behavioral performance and cognition, and UA also prevents the cell death of dopaminergic neurons, which is induced by modulation of neuroinflammation and oxidative stress [101]. UA treatment enhances Nrf2-responsive gene expression, including GCLC, HO-1, and NQO1, increases SOD, catalase (CAT), and GSH levels, and reduces MDA in the substantia nigra [101] (Table 3).
UA also has an indirect neuroprotective activity by increasing GSH levels in astrocytes. UA treatment activates Nrf2 and leads to upregulation of gene expression of the GCLM, which enhances GSH production in astrocytes [37] (Table 3). This increase in astrocytic GSH levels can enhance GSH release into extracellular space, and the released GSH can be converted to Cys, which is taken up by neurons [37].
Nrf2 activation is another indirect mechanism by which purine derivatives provide neuroprotection. Nrf2 activation was observed when caffeine was treated as a neuroprotective agent in lipopolysaccharide (LPS)-induced oxidative stress model mice [134] and cadmium-induced cognitive deficit model mice [135] (Table 3). Nrf2 activators such as tert-butylhydroquinone (t-BHQ) and L-sulforaphane can induce EAAC1 expression by the Nrf2/ARE pathway in C6 glioma cells [141]. In vivo treatment of t-BHQ upregulates EAAC1 expression by Nrf2 activation and increases neuronal GSH levels in the mouse striatum [141]. T-BHQ also upregulates xCT expression in astrocytes in vitro [142]. Although caffeine-induced EAAC1 expression via A2AAR antagonism is observed in the developing retina [143], it is still unclear whether caffeine-induced Nrf2 activation upregulates EAAC1 or xCT and leads to increasing GSH synthesis.
Our previous study showed a novel mechanism by which treatment with caffeine (10 to 100 µM) and UA (1 to 10 µM) provided neuroprotection by increasing GSH levels in the brain. The upregulation of GSH synthesis is mediated by increasing Cys uptake via EAAC1 in hippocampal slices [20]. In addition, paraxanthine (1,7-dimethylxanthine), a major metabolite of caffeine, increases Cys uptake in hippocampal slices whereas other purine metabolites (theophylline, theobromine, 1-MX, 3-MX, and 1,7-methyluricacid) do not [21]. Paraxanthine (10 to 100 µM) increases Cys uptake in human neuroblastoma SH-SY5Y cells via EAAC1 transport activity, and paraxanthine also exhibits neuroprotective activity in SH-SY5Y cells [21]. The upregulation of GSH levels is independent of both A1AR and A2AAR antagonisms, because an A2AAR antagonist, SCH58261, did not increase GSH levels in hippocampal slices [20], and because UA exhibits no activity on A1AR (IC50 > 5300 µM, Table 2) [68].
In astrocytes, guanosine upregulates GCL, glutamine synthetase (GS), and GSH reductase (GR), resulting in elevated GSH levels [136,137] (Table 3). The increase in GSH levels in astrocytes may protect neurons against oxidative stress by promoting GSH supply to neurons. Thus, GSH-regulated gene expression is one of the neuroprotective mechanisms of UA, caffeine, and guanosine alike.
Finally, guanosine is a nucleotide metabolite that acts as an efficient neuromodulator in the brain, and its extracellular role has recently been clarified [4,137,144]. Extracellular guanosine is found to be neuroprotective and active in neurological regeneration in response to brain ischemia and trauma [4]. The mechanisms of guanosine-induced neuroprotection involve A1AR and A2AAR (of which it is a weak agonist [4]), Kir 4.1 potassium channels (Kir 4.1), and the excitatory amino acid transporter, glutamate transporter 1 (GLT-1). In neurons and astrocytes, guanosine activates pro-survival pathways such as Phosphoinositide 3-Kinase (PI3K), protein kinase B (Akt), and extracellular-signal regulated kinase kinase (MEK)/extracellular-signal regulated kinase (ERK) via adenosine receptors [138,145]. Guanosine also exerts neuroprotective activities by upregulations of PI3K/Akt and MEK/ERK signals in rat cortical astrocytes [138], and upregulation of PI3K/Akt/glycogen synthase kinase 3 (GSK3) signal in rat hippocampus slice [139] (Table 3). In vivo treatment of guanosine treatment improves behavioral performance and reduces mitochondrial dysfunction in the penumbra area in ischemia model rat [146].

3.3.2. Neuroprotection by Adenosine Receptor Modulation

Adenosine is another purine that has neuroprotective activities. Since the 1940s, adenosine has been used clinically for cardiac protection and vasodilation [147]. Adenosine exhibits both neurostimulative and neuroprotective activities. Its neuroprotective activity is typified as repression of excess neuronal activation (category (i) in Table 1). Extracellular adenosine at concentrations of 0.05 to 0.2 µM is sufficient to modulate synaptic functions [147]. The actions of adenosine in the brain are mediated by A1AR (IC50 = 0.070 µM), A2AAR (IC50 = 0.150 µM), adenosine A2B receptor (A2BAR, Gene ID: 136) (IC50 = 15.4 µM), and adenosine A3 receptor (A3AR, Gene ID: 140) (IC50 = 6.5 µM), which are variously activated, depending on adenosine concentrations [67,148,149] (Table 2). The expression of subtypes of adenosine receptors varies according to cell type. A1AR is the most abundant in the brain. Adenosine reduces neuronal excitability and firing rate via A1AR stimulation [150,151,152]. A1AR stimulation by adenosine induces neuroprotection in cultured neurons and ischemia model of rats [2,140] (Figure 3, Table 3). In vivo studies have shown that neuroprotection was induced by A1AR stimulation in animal models of ischemia [152,153]. A1AR stimulation inhibits Ca2+ entry into the presynaptic terminal, resulting in a reduction in neurotransmitter release [154,155,156] and hyperpolarizes postsynaptic neurons [157,158]. In contrast, A2AAR stimulation in the brain induces neuronal excitability and synaptic transmission [147,159]. Specific A2AAR antagonists such as KW-6002 (istradefylline), which is xanthine-based compound, protect nigral dopaminergic cells from damage induced by 6-OHDA in rats or by MPTP in mice [160]. Genetic depletion of A2AAR lowered the dopaminergic neurotoxicity in PD model animals [62]. These results indicate neuroprotective activities of both A1AR agonism and A2AAR antagonism [3].
Caffeine is another purine that offers category (i) neuroprotection through the A2AAR receptor, as well as its previously described category (ii) role of increasing intracellular GSH levels via Cys uptake. Methylxanthines such as caffeine and theophylline are known for their bronchoprotective effects. The molecular mechanisms include adenosine receptor antagonism (Caffeine; IC50 = 107 µM) and phosphodiesterase inhibition (Caffeine; IC50 = 400 µM) [69] (Table 2). A 30 mg/kg intraperitoneal injection of caffeine provides neuroprotective effects (Table 3). Since the serum caffeine concentration is 116 µM 60 min after the injection [161], it is enough concentration to antagonize adenosine receptors but not to affect phosphodiesterase activity. Of course, caffeine also has motor stimulant, psychostimulant, arousal, anti-inflammatory, anti-oxidative, and neuroprotective effects [69,162]. A1AR antagonism is involved in the stimulus effects of caffeine [163]. Of these, the arousal effect of caffeine (15 mg/kg, i.p.) is dependent on the A2AAR antagonism [164], and the anti-inflammatory effect of caffeine (100 µM) is mediated by both A1AR and A2AAR [165,166]. The locomotor stimulatory effect of high doses of caffeine cannot be attributed to the modulation of either the A1AR or the A2AAR, suggesting that this effect is independent of AR activity. It is accepted that the neuroprotective activities of caffeine are mediated by A2AAR inhibition [3]. Caffeine, paraxanthine, or theophylline (10 or 30 mg/kg, i.p.) prevents neuronal cell loss in MPTP-treated PD model animals [5,6] (Table 3). In fat-enriched diet-induced cognitive deficits rats, chronic treatment with theobromine (30 mg/day) improved cognitive functions and amyloid-beta protein (Aβ) and interleukin-1β (IL-1β) levels in the brain [7] (Table 3). Chronic treatment with caffeine resulted in tolerance to its motor stimulation activity, whereas caffeine-induced neuroprotection does not diminish with exposure [167]. Mechanisms other than adenosine receptors may also be involved in the neuroprotective effects of caffeine.

3.3.3. Neuroprotection by Other Mechanisms

There are a few other, less pervasive mechanisms responsible for neuroprotective activities of purines. UA promotes pro-survival pathways such as activating PI3K/Akt, and MEK/ERK in the brain. For example, UA treatment (200 µM UA) prevents dopamine (DA) neuron loss and behavioral deficits in 6-OHDA treated rat by recovery of Akt/GSK3β signaling [8] (Table 3). A PI3K inhibitor was shown to interfere with UA-induced neuroprotection and regulations on Akt/GSK3β signaling in 6-OHDA-treated SH-SY5Y cells [8]. Thus, PI3K/Akt/GSK3β signaling can be involved in the neuroprotective activities of UA.
Ya et al. [9] have shown that UA treatment (UA 16 mg/kg i.v.) reduces focal cerebral ischemic reperfusion-induced oxidative stress, preventing neuronal damage. The mechanisms underlying this neuroprotective activity of UA are the upregulations of both brain-derived neurotrophic factor (BDNF) and nerve growth factor (NGF) via the Nrf2 signaling pathway.
Paraxanthine at a concentration of 800 µM has a neuroprotective activity for dopaminergic neurons [84]. Its neuroprotective mechanism is attributed to a cytosolic calcium release from the endoplasmic reticulum via the activation of ryanodine receptor channels. However, this activity of paraxanthine is not mediated by antagonizing adenosine receptors or by elevation of intracellular cAMP (phosphodiesterase inhibition).
ATP itself has no neuroprotective activity corresponding to categories (i) and (ii) in Table 1. However, ATP increases the expression of the astrocyte glutamate transporter, GLT-1, by P2Y (ATP receptor) stimulation, which activates the ERK/nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) signaling pathway. Since upregulation of GLT-1 in astrocyte promotes the removal of excess glutamate from extra-synaptic space, ATP may be involved in this category (i) neuroprotective mechanism. Thus, extracellular purines such as adenosine and ATP regulate cellular signaling via their specific receptors.
As described above, the neuroprotective mechanisms of purine derivatives are mainly due to adenosine receptor modulation and upregulation of the synthesis of proteins that are involved in increasing antioxidative activity or activating pro-survival pathways in cells. Many receptor activities develop tolerance to their antagonists [167], i.e., caffeine-induced arousal, psychostimulant, and motor stimulant activities develop tolerance to caffeine, but the neuroprotective activity does not. In addition to antagonizing A2AAR, increasing GSH levels via Cys uptake may play an important role in the neuroprotective activities of purine derivatives.

4. Conclusions

Several purine derivatives exhibit neuroprotective activity mediated by increasing neuronal GSH levels, which can be due to not only induction of GSH synthesis-related enzymes, but also the upregulation of GSH levels explained by promoting EAAC1-mediated Cys uptake, which is a commonly seen after treatment with caffeine, UA, or paraxanthine.
The upregulation of GSH levels may enhance antioxidative activity in neurons and appears to be effective in preventing the incidence and progression of neurodegenerative disease. Further studies to investigate how purine derivatives increase GSH levels by promoting EAAC1-mediated Cys uptake would provide valuable insights into neuroprotection by purine derivative treatment. These studies of purine derivative biochemistry could lead to novel preventions and treatments for the growing burden of neurodegenerative disease in aging societies worldwide.

Author Contributions

Conceptualization: N.M. and K.A.; literature searches and manuscript writing: N.M.; supervision: K.A. All authors have read and agreed to the published version of the manuscript.

Funding

Author N. Matsumura has received research support from the Urakami Foundation for Food and Food Culture Promotion, Japan (grant no. R01205).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

5′NT, CD735′-nucleotidase
6-OHDA6-hydroxydopamine
8-OHdG8-hydroxyl-2′-deoxyguanosine
A1ARadenosine A1 receptor
A2AARadenosine A2A receptor
A2BARadenosine A2B receptor
A3ARadenosine A3 receptor
ADAlzheimer disease
ADAadenosine deaminase
Aktprotein kinase B
ApNaminopeptidase N
amyloid-beta protein
BBBblood–brain barrier
BDNFbrain-derived neurotrophic factor
cAMPcyclic AMP
CATcatalase
CNScentral nervous system
CYP cytochrome P450.
CYP1A2cytochrome P450 family 1 subfamily A member 2
CysCysteine
CysCysCystine
CysGlyCysteinylglycine
DADopamine
EAAC1Excitatory amino acid carrier protein 1
ERKExtracellular-signal regulated kinase
GCLγ-glutamyl cysteine ligase
GCLCGCL catalytic subunit
GCLMGCL modifier subunit
GDAGuanine deaminase
GLASTGlutamate-aspartate transporter
GlnGlutamine
GLT-1Glutamate transporter 1
GluGlutamate
GluCysglutamylcysteine
GluRglutamate transporter
Glyglycine
GPxglutathione peroxidase
GRGSH reductase
GSglutamine synthetase
GSHglutathione
GSK3glycogen synthase kinase 3
GSK3βglycogen synthase kinase 3 beta
GSSGSH synthetase
GSSGglutathione disulfide
GSTglutathione S-transferase
HO-1heme oxygenase-1
IL-1βinterleukin-1β
i.v.intravenously injection
Kir 4.1Kir 4.1 potassium channel
LDHlactate dehydrogenase
LPSlipopolysaccharide
MDAmalondialdehyde
MEKextracellular-signal regulated kinase kinase
MPP+1-methyl-4-phenylpyridinium
MPTP1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine
MRP1multidrug resistance protein 1
MRPsmultidrug resistance proteins
MXmethylxanthine
NF-κBnuclear factor kappa-light-chain-enhancer of activated B cells
NGFnerve growth factor
NQO1NAD(P)H quinone oxidoreductase 1
Nrf2nuclear factor erythroid-2-related factor 2
NTPDases, CD39nucleoside triphosphate diphospho-hydrolase
PDParkinson’s disease
PI3KPhosphoinositide 3-Kinase
p-JNKphospho-c-Jun n-terminal kinase
p-NF-κBphosphor-NF-κB
PNPpurine nucleoside phosphorylase
PXparaxanthine
ROSreactive oxygen species
SIN-13-morpholinosydnonimine
SODsuperoxide dismutase
TBtheobromine
t-BHQtert-butylhydroquinone
TPtheophylline
TUNELterminal deoxynucleotidyl transferase-mediated dNTP nick end labeling
UAuric acid
xCTcystine/glutamate antiporter
XOxanthine oxidase
γ-GTγ-glutamyl transpeptidase

References

  1. Stone, T.W. Purines and neuroprotection. Adv. Exp. Med. Biol. 2002, 513, 249–280. [Google Scholar] [CrossRef] [PubMed]
  2. Jiang, N.; Kowaluk, E.A.; Lee, C.H.; Mazdiyasni, H.; Chopp, M. Adenosine kinase inhibition protects brain against transient focal ischemia in rats. Eur. J. Pharmacol. 1997, 320, 131–137. [Google Scholar] [CrossRef] [PubMed]
  3. Cunha, R.A. Neuroprotection by adenosine in the brain: From A(1) receptor activation to A (2A) receptor blockade. Purinergic Signal. 2005, 1, 111–134. [Google Scholar] [CrossRef] [PubMed]
  4. Chojnowski, K.; Opielka, M.; Nazar, W.; Kowianski, P.; Smolenski, R.T. Neuroprotective Effects of Guanosine in Ischemic Stroke-Small Steps towards Effective Therapy. Int. J. Mol. Sci. 2021, 22, 6898. [Google Scholar] [CrossRef]
  5. Chen, J.-F.; Xu, K.; Petzer, J.P.; Staal, R.; Xu, Y.-H.; Beilstein, M.; Sonsalla, P.K.; Castagnoli, K.; Castagnoli, N.; Schwarzschild, M.A. Neuroprotection by caffeine and A(2A) adenosine receptor inactivation in a model of Parkinson’s disease. J. Neurosci. 2001, 21, RC143. [Google Scholar] [CrossRef]
  6. Xu, K.; Xu, Y.H.; Chen, J.F.; Schwarzschild, M.A. Neuroprotection by caffeine: Time course and role of its metabolites in the MPTP model of Parkinson’s disease. Neuroscience 2010, 167, 475–481. [Google Scholar] [CrossRef]
  7. Mendiola-Precoma, J.; Padilla, K.; Rodriguez-Cruz, A.; Berumen, L.C.; Miledi, R.; Garcia-Alcocer, G. Theobromine-Induced Changes in A1 Purinergic Receptor Gene Expression and Distribution in a Rat Brain Alzheimer’s Disease Model. J. Alzheimers Dis. 2017, 55, 1273–1283. [Google Scholar] [CrossRef]
  8. Gong, L.; Zhang, Q.-L.; Zhang, N.; Hua, W.-Y.; Huang, Y.-X.; Di, P.-W.; Huang, T.; Xu, X.-S.; Liu, C.-F.; Hu, L.-F.; et al. Neuroprotection by urate on 6-OHDA-lesioned rat model of Parkinson’s disease: Linking to Akt/GSK3beta signaling pathway. J. Neurochem. 2012, 123, 876–885. [Google Scholar] [CrossRef]
  9. Ya, B.-L.; Liu, Q.; Li, H.-F.; Cheng, H.-J.; Yu, T.; Chen, L.; Wang, Y.; Yuan, L.-L.; Li, W.-J.; Liu, W.-Y.; et al. Uric Acid Protects against Focal Cerebral Ischemia/Reperfusion-Induced Oxidative Stress via Activating Nrf2 and Regulating Neurotrophic Factor Expression. Oxid. Med. Cell. Longev. 2018, 2018, 6069150. [Google Scholar] [CrossRef]
  10. Frei, B.; Stocker, R.; Ames, B.N. Antioxidant defenses and lipid peroxidation in human blood plasma. Proc. Natl. Acad. Sci. USA 1988, 85, 9748–9752. [Google Scholar] [CrossRef]
  11. Frei, B.; England, L.; Ames, B.N. Ascorbate is an outstanding antioxidant in human blood plasma. Proc. Natl. Acad. Sci. USA 1989, 86, 6377–6381. [Google Scholar] [CrossRef] [PubMed]
  12. Santos, C.; Lunet, N.; Azevedo, A.; de Mendonca, A.; Ritchie, K.; Barros, H. Caffeine intake is associated with a lower risk of cognitive decline: A cohort study from Portugal. J. Alzheimers Dis. 2010, 20 (Suppl. 1), S175–S185. [Google Scholar] [CrossRef] [PubMed]
  13. Dong, X.; Li, S.; Sun, J.; Li, Y.; Zhang, D. Association of Coffee, Decaffeinated Coffee and Caffeine Intake from Coffee with Cognitive Performance in Older Adults: National Health and Nutrition Examination Survey (NHANES) 2011–2014. Nutrients 2020, 12, 840. [Google Scholar] [CrossRef] [PubMed]
  14. Irizarry, M.C.; Raman, R.; Schwarzschild, M.A.; Becerra, L.M.; Thomas, R.G.; Peterson, R.C.; Ascherio, A.; Aisen, P.S. Plasma urate and progression of mild cognitive impairment. Neurodegener Dis. 2009, 6, 23–28. [Google Scholar] [CrossRef]
  15. Scheepers, L.; Jacobsson, L.T.H.; Kern, S.; Johansson, L.; Dehlin, M.; Skoog, I. Urate and risk of Alzheimer’s disease and vascular dementia: A population-based study. Alzheimer’s Dement. 2019, 15, 754–763. [Google Scholar] [CrossRef]
  16. Ascherio, A.; Zhang, S.M.; Hernan, M.A.; Kawachi, I.; Colditz, G.A.; Speizer, F.E.; Willett, W.C. Prospective study of caffeine consumption and risk of Parkinson’s disease in men and women. Ann. Neurol. 2001, 50, 56–63. [Google Scholar] [CrossRef]
  17. Ross, G.W.; Abbott, R.D.; Petrovitch, H.; Morens, D.M.; Grandinetti, A.; Tung, K.-H.; Tanner, C.M.; Masaki, K.H.; Blanchette, P.L.; Curb, J.D.; et al. Association of coffee and caffeine intake with the risk of Parkinson disease. JAMA 2000, 283, 2674–2679. [Google Scholar] [CrossRef]
  18. Cortese, M.; Riise, T.; Engeland, A.; Ascherio, A.; Bjornevik, K. Urate and the risk of Parkinson’s disease in men and women. Parkinsonism Relat. Disord. 2018, 52, 76–82. [Google Scholar] [CrossRef]
  19. Bakshi, R.; Macklin, E.A.; Hung, A.Y.; Hayes, M.T.; Hyman, B.T.; Wills, A.-M.; Gomperts, S.N.; Growdon, J.H.; Ascherio, A.; Scherzer, C.R.; et al. Associations of Lower Caffeine Intake and Plasma Urate Levels with Idiopathic Parkinson’s Disease in the Harvard Biomarkers Study. J. Parkinsons Dis. 2020, 10, 505–510. [Google Scholar] [CrossRef]
  20. Aoyama, K.; Matsumura, N.; Watabe, M.; Wang, F.; Kikuchi-Utsumi, K.; Nakaki, T. Caffeine and uric acid mediate glutathione synthesis for neuroprotection. Neuroscience 2011, 181, 206–215. [Google Scholar] [CrossRef]
  21. Matsumura, N.; Kinoshita, C.; Bhadhprasit, W.; Nakaki, T.; Aoyama, K. A purine derivative, paraxanthine, promotes cysteine uptake for glutathione synthesis. J. Pharmacol. Sci. 2023, 151, 37–45. [Google Scholar] [CrossRef] [PubMed]
  22. Dringen, R. Metabolism and functions of glutathione in brain. Prog. Neurobiol. 2000, 62, 649–671. [Google Scholar] [CrossRef] [PubMed]
  23. Aoyama, K. Glutathione in the Brain. Int. J. Mol. Sci. 2021, 22, 5010. [Google Scholar] [CrossRef] [PubMed]
  24. Muir, K.W.; Lees, K.R. Clinical experience with excitatory amino acid antagonist drugs. Stroke 1995, 26, 503–513. [Google Scholar] [CrossRef] [PubMed]
  25. Wu, Q.J.; Tymianski, M. Targeting NMDA receptors in stroke: New hope in neuroprotection. Mol. Brain 2018, 11, 15. [Google Scholar] [CrossRef]
  26. Wang, S.J.; Sihra, T.S.; Gean, P.W. Lamotrigine inhibition of glutamate release from isolated cerebrocortical nerve terminals (synaptosomes) by suppression of voltage-activated calcium channel activity. Neuroreport 2001, 12, 2255–2258. [Google Scholar] [CrossRef]
  27. Cheah, B.C.; Vucic, S.; Krishnan, A.V.; Kiernan, M.C. Riluzole, neuroprotection and amyotrophic lateral sclerosis. Curr. Med. Chem. 2010, 17, 1942–1992. [Google Scholar] [CrossRef]
  28. Bélanger, M.; Magistretti, P.J. The role of astroglia in neuroprotection. Dialogues Clin. Neurosci. 2009, 11, 281–295. [Google Scholar] [CrossRef]
  29. Pajarillo, E.; Rizor, A.; Lee, J.; Aschner, M.; Lee, E. The role of astrocytic glutamate transporters GLT-1 and GLAST in neurological disorders: Potential targets for neurotherapeutics. Neuropharmacology 2019, 161, 107559. [Google Scholar] [CrossRef]
  30. Gipp, J.J.; Bailey, H.H.; Mulcahy, R.T. Cloning and sequencing of the cDNA for the light subunit of human liver gamma-glutamylcysteine synthetase and relative mRNA levels for heavy and light subunits in human normal tissues. Biochem. Biophys. Res. Commun. 1995, 206, 584–589. [Google Scholar] [CrossRef]
  31. Gipp, J.J.; Chang, C.; Mulcahy, R.T. Cloning and nucleotide sequence of a full-length cDNA for human liver gamma-glutamylcysteine synthetase. Biochem. Biophys. Res. Commun. 1992, 185, 29–35. [Google Scholar] [CrossRef] [PubMed]
  32. Lu, S.C. Glutathione synthesis. Biochim. Biophys. Acta 2013, 1830, 3143–3153. [Google Scholar] [CrossRef]
  33. Oppenheimer, L.; Wellner, V.P.; Griffith, O.W.; Meister, A. Glutathione synthetase. Purification from rat kidney and mapping of the substrate binding sites. J. Biol. Chem. 1979, 254, 5184–5190. [Google Scholar] [PubMed]
  34. Rothstein, J.D.; Martin, L.; Levey, A.I.; Dykes-Hoberg, M.; Jin, L.; Wu, D.; Nash, N.; Kuncl, R.W. Localization of neuronal and glial glutamate transporters. Neuron 1994, 13, 713–725. [Google Scholar] [CrossRef] [PubMed]
  35. Zerangue, N.; Kavanaugh, M.P. Interaction of L-cysteine with a human excitatory amino acid transporter. J. Physiol. 1996, 493 Pt 2, 419–423. [Google Scholar] [CrossRef]
  36. Aoyama, K.; Suh, S.W.; Hamby, A.M.; Liu, J.; Chan, W.Y.; Chen, Y.; Swanson, R.A. Neuronal glutathione deficiency and age-dependent neurodegeneration in the EAAC1 deficient mouse. Nat. Neurosci. 2006, 9, 119–126. [Google Scholar] [CrossRef]
  37. Bakshi, R.; Zhang, H.; Logan, R.; Joshi, I.; Xu, Y.; Chen, X.; Schwarzschild, M.A. Neuroprotective effects of urate are mediated by augmenting astrocytic glutathione synthesis and release. Neurobiol. Dis. 2015, 82, 574–579. [Google Scholar] [CrossRef]
  38. Sato, H.; Tamba, M.; Okuno, S.; Sato, K.; Keino-Masu, K.; Masu, M.; Bannai, S. Distribution of cystine/glutamate exchange transporter, system x(c)-, in the mouse brain. J. Neurosci. 2002, 22, 8028–8033. [Google Scholar] [CrossRef]
  39. Hirrlinger, J.; Schulz, J.B.; Dringen, R. Glutathione release from cultured brain cells: Multidrug resistance protein 1 mediates the release of GSH from rat astroglial cells. J. Neurosci. Res. 2002, 69, 318–326. [Google Scholar] [CrossRef]
  40. Rebec, G.V. Dysregulation of corticostriatal ascorbate release and glutamate uptake in transgenic models of Huntington’s disease. Antioxid Redox Signal. 2013, 19, 2115–2128. [Google Scholar] [CrossRef]
  41. Padayatty, S.J.; Katz, A.; Wang, Y.; Eck, P.; Kwon, O.; Lee, J.-H.; Chen, S.; Corpe, C.; Dutta, A.; Dutta, S.K.; et al. Vitamin C as an antioxidant: Evaluation of its role in disease prevention. J. Am. Coll Nutr. 2003, 22, 18–35. [Google Scholar] [CrossRef] [PubMed]
  42. Harrison, F.E.; Bowman, G.L.; Polidori, M.C. Ascorbic acid and the brain: Rationale for the use against cognitive decline. Nutrients 2014, 6, 1752–1781. [Google Scholar] [CrossRef] [PubMed]
  43. Schirinzi, T.; Martella, G.; Imbriani, P.; Di Lazzaro, G.; Franco, D.; Colona, V.L.; Alwardat, M.; Salimei, P.S.; Mercuri, N.B.; Pierantozzi, M.; et al. Dietary Vitamin E as a Protective Factor for Parkinson’s Disease: Clinical and Experimental Evidence. Front. Neurol. 2019, 10, 148. [Google Scholar] [CrossRef] [PubMed]
  44. Ghosh, C.; Hossain, M.; Solanki, J.; Dadas, A.; Marchi, N.; Janigro, D. Pathophysiological implications of neurovascular P450 in brain disorders. Drug Discov. Today 2016, 21, 1609–1619. [Google Scholar] [CrossRef]
  45. Mann, A.; Miksys, S.L.; Gaedigk, A.; Kish, S.J.; Mash, D.C.; Tyndale, R.F. The neuroprotective enzyme CYP2D6 increases in the brain with age and is lower in Parkinson’s disease patients. Neurobiol. Aging 2012, 33, 2160–2171. [Google Scholar] [CrossRef]
  46. Zhang, M.; An, C.; Gao, Y.; Leak, R.K.; Chen, J.; Zhang, F. Emerging roles of Nrf2 and phase II antioxidant enzymes in neuroprotection. Prog. Neurobiol. 2013, 100, 30–47. [Google Scholar] [CrossRef]
  47. Gordillo, G.M.; Biswas, A.; Khanna, S.; Spieldenner, J.M.; Pan, X.; Sen, C.K. Multidrug Resistance-associated Protein-1 (MRP-1)-dependent Glutathione Disulfide (GSSG) Efflux as a Critical Survival Factor for Oxidant-enriched Tumorigenic Endothelial Cells. J. Biol. Chem. 2016, 291, 10089–10103. [Google Scholar] [CrossRef]
  48. Rush, T.; Liu, X.; Nowakowski, A.B.; Petering, D.H.; Lobner, D. Glutathione-mediated neuroprotection against methylmercury neurotoxicity in cortical culture is dependent on MRP1. Neurotoxicology 2012, 33, 476–481. [Google Scholar] [CrossRef]
  49. Aoyama, K.; Nakaki, T. Glutathione in Cellular Redox Homeostasis: Association with the Excitatory Amino Acid Carrier 1 (EAAC1). Molecules 2015, 20, 8742–8758. [Google Scholar] [CrossRef]
  50. Bendahan, A.; Armon, A.; Madani, N.; Kavanaugh, M.P.; Kanner, B.I. Arginine 447 plays a pivotal role in substrate interactions in a neuronal glutamate transporter. J. Biol. Chem. 2000, 275, 37436–37442. [Google Scholar] [CrossRef]
  51. Dringen, R.; Pfeiffer, B.; Hamprecht, B. Synthesis of the antioxidant glutathione in neurons: Supply by astrocytes of CysGly as precursor for neuronal glutathione. J. Neurosci. 1999, 19, 562–569. [Google Scholar] [CrossRef] [PubMed]
  52. Minich, T.; Riemer, J.; Schulz, J.B.; Wielinga, P.; Wijnholds, J.; Dringen, R. The multidrug resistance protein 1 (Mrp1), but not Mrp5, mediates export of glutathione and glutathione disulfide from brain astrocytes. J. Neurochem. 2006, 97, 373–384. [Google Scholar] [CrossRef] [PubMed]
  53. Resende, R.; Moreira, P.I.; Proenca, T.; Deshpande, A.; Busciglio, J.; Pereira, C.; Oliveira, C.R. Brain oxidative stress in a triple-transgenic mouse model of Alzheimer disease. Free Radic. Biol. Med. 2008, 44, 2051–2057. [Google Scholar] [CrossRef]
  54. Jenner, P. Oxidative damage in neurodegenerative disease. Lancet 1994, 344, 796–798. [Google Scholar] [CrossRef]
  55. Bettens, K.; Sleegers, K.; Van Broeckhoven, C. Genetic insights in Alzheimer’s disease. Lancet Neurol. 2013, 12, 92–104. [Google Scholar] [CrossRef] [PubMed]
  56. Trinh, J.; Farrer, M. Advances in the genetics of Parkinson disease. Nat. Rev. Neurol. 2013, 9, 445–454. [Google Scholar] [CrossRef]
  57. Lindsay, J.; Laurin, D.; Verreault, R.; Hebert, R.; Helliwell, B.; Hill, G.B.; McDowell, I. Risk factors for Alzheimer’s disease: A prospective analysis from the Canadian Study of Health and Aging. Am. J. Epidemiol. 2002, 156, 445–453. [Google Scholar] [CrossRef]
  58. Delamarre, A.; Meissner, W.G. Epidemiology, environmental risk factors and genetics of Parkinson’s disease. Presse Med. 2017, 46 Pt 1, 175–181. [Google Scholar] [CrossRef]
  59. Langston, J.W. The etiology of Parkinson’s disease with emphasis on the MPTP story. Neurology 1996, 47 (Suppl. 3), S153–S160. [Google Scholar] [CrossRef]
  60. Fujimaki, M.; Saiki, S.; Li, Y.; Kaga, N.; Taka, H.; Hatano, T.; Ishikawa, K.I.; Oji, Y.; Mori, A.; Okuzumi, A.; et al. Serum caffeine and metabolites are reliable biomarkers of early Parkinson disease. Neurology 2018, 90, e404–e411. [Google Scholar] [CrossRef]
  61. Xu, K.; Di Luca, D.G.; Orru, M.; Xu, Y.; Chen, J.F.; Schwarzschild, M.A. Neuroprotection by caffeine in the MPTP model of parkinson’s disease and its dependence on adenosine A2A receptors. Neuroscience 2016, 322, 129–137. [Google Scholar] [CrossRef] [PubMed]
  62. Magkos, F.; Kavouras, S.A. Caffeine use in sports, pharmacokinetics in man, and cellular mechanisms of action. Crit. Rev. Food Sci. Nutr. 2005, 45, 535–562. [Google Scholar] [CrossRef] [PubMed]
  63. Guerreiro, S.; Ponceau, A.; Toulorge, D.; Martin, E.; Alvarez-Fischer, D.; Hirsch, E.C.; Michel, P.P. Protection of midbrain dopaminergic neurons by the end-product of purine metabolism uric acid: Potentiation by low-level depolarization. J. Neurochem. 2009, 109, 1118–1128. [Google Scholar] [CrossRef] [PubMed]
  64. Nishida, Y. Inhibition of lipid peroxidation by methylated analogues of uric acid. J. Pharm. Pharmacol. 1991, 43, 885–887. [Google Scholar] [CrossRef] [PubMed]
  65. Fredholm, B.B.; Lindstrom, K. Autoradiographic comparison of the potency of several structurally unrelated adenosine receptor antagonists at adenosine A1 and A(2A) receptors. Eur. J. Pharmacol. 1999, 380, 197–202. [Google Scholar] [CrossRef]
  66. Klotz, K.N. Adenosine receptors and their ligands. Naunyn Schmiedebergs Arch. Pharmacol. 2000, 362, 382–391. [Google Scholar] [CrossRef]
  67. Muller, C.E.; Scior, T. Adenosine receptors and their modulators. Pharm. Acta Helv. 1993, 68, 77–111. [Google Scholar] [CrossRef]
  68. Hunter, R.E.; Barrera, C.M.; Dohanich, G.P.; Dunlap, W.P. Effects of uric acid and caffeine on A1 adenosine receptor binding in developing rat brain. Pharmacol. Biochem. Behav. 1990, 35, 791–795. [Google Scholar] [CrossRef]
  69. Fredholm, B.B.; Battig, K.; Holmen, J.; Nehlig, A.; Zvartau, E.E. Actions of caffeine in the brain with special reference to factors that contribute to its widespread use. Pharmacol. Rev. 1999, 51, 83–133. [Google Scholar]
  70. Jimenez-Jimenez, F.J.; Mateo, D.; Gimenez-Roldan, S. Premorbid smoking, alcohol consumption, and coffee drinking habits in Parkinson’s disease: A case-control study. Mov. Disord. 1992, 7, 339–344. [Google Scholar] [CrossRef]
  71. Grandinetti, A.; Morens, D.M.; Reed, D.; MacEachern, D. Prospective study of cigarette smoking and the risk of developing idiopathic Parkinson’s disease. Am. J. Epidemiol. 1994, 139, 1129–1138. [Google Scholar] [CrossRef] [PubMed]
  72. Morano, A.; Jimenez-Jimenez, F.J.; Molina, J.A.; Antolin, M.A. Risk-factors for Parkinson’s disease: Case-control study in the province of Caceres, Spain. Acta Neurol. Scand 1994, 89, 164–170. [Google Scholar] [CrossRef]
  73. Hellenbrand, W.; Seidler, A.; Boeing, H.; Robra, B.-P.; Vieregge, P.; Nischan, P.; Joerg, J.; Oertel, W.H.; Schneider, E.; Ulm, G. Diet and Parkinson’s disease. I: A possible role for the past intake of specific foods and food groups. Results from a self-administered food-frequency questionnaire in a case-control study. Neurology 1996, 47, 636–643. [Google Scholar] [CrossRef] [PubMed]
  74. Chan, D.K.Y.; Woo, J.; Ho, S.C.; Pang, C.P.; Law, L.K.; Ng, P.W.; Hung, W.T.; Kwok, T.; Hui, E.; Orr, K.; et al. Genetic and environmental risk factors for Parkinson’s disease in a Chinese population. J. Neurol. Neurosurg. Psychiatry 1998, 65, 781–784. [Google Scholar] [CrossRef] [PubMed]
  75. Fall, P.A.; Fredrikson, M.; Axelson, O.; Granerus, A.K. Nutritional and occupational factors influencing the risk of Parkinson’s disease: A case-control study in southeastern Sweden. Mov. Disord. 1999, 14, 28–37. [Google Scholar] [CrossRef]
  76. Benedetti, M.D.; Bower, J.H.; Maraganore, D.M.; McDonnell, S.K.; Peterson, B.J.; Ahlskog, J.E.; Schaid, D.J.; Rocca, W.A. Smoking, alcohol, and coffee consumption preceding Parkinson’s disease: A case-control study. Neurology 2000, 55, 1350–1358. [Google Scholar] [CrossRef]
  77. Checkoway, H.; Powers, K.; Smith-Weller, T.; Franklin, G.M.; Longstreth, W.T., Jr.; Swanson, P.D. Parkinson’s disease risks associated with cigarette smoking, alcohol consumption, and caffeine intake. Am. J. Epidemiol. 2002, 155, 732–738. [Google Scholar] [CrossRef]
  78. Ragonese, P.; Salemi, G.; Morgante, L.; Aridon, P.; Epifanio, A.; Buffa, D.; Scoppa, F.; Savettieri, G. A case-control study on cigarette, alcohol, and coffee consumption preceding Parkinson’s disease. Neuroepidemiology 2003, 22, 297–304. [Google Scholar] [CrossRef]
  79. Tan, E.-K.; Tan, C.; Fook-Chong, S.; Lum, S.; Chai, A.; Chung, H.; Shen, H.; Zhao, Y.; Teoh, M.; Yih, Y.; et al. Dose-dependent protective effect of coffee, tea, and smoking in Parkinson’s disease: A study in ethnic Chinese. J. Neurol. Sci. 2003, 216, 163–167. [Google Scholar] [CrossRef]
  80. Nehlig, A. Interindividual Differences in Caffeine Metabolism and Factors Driving Caffeine Consumption. Pharmacol. Rev. 2018, 70, 384–411. [Google Scholar] [CrossRef]
  81. Sachse, C.; Brockmoller, J.; Bauer, S.; Roots, I. Functional significance of a C-->A polymorphism in intron 1 of the cytochrome P450 CYP1A2 gene tested with caffeine. Br. J. Clin. Pharmacol. 1999, 47, 445–449. [Google Scholar] [CrossRef]
  82. Tan, E.K.; Chua, E.; Fook-Chong, S.M.; Teo, Y.Y.; Yuen, Y.; Tan, L.; Zhao, Y. Association between caffeine intake and risk of Parkinson’s disease among fast and slow metabolizers. Pharmacogenet Genom. 2007, 17, 1001–1005. [Google Scholar] [CrossRef] [PubMed]
  83. Leodori, G.; De Bartolo, M.I.; Belvisi, D.; Ciogli, A.; Fabbrini, A.; Costanzo, M.; Manetto, S.; Conte, A.; Villani, C.; Fabbrini, G.; et al. Salivary caffeine in Parkinson’s disease. Sci. Rep. 2021, 11, 9823. [Google Scholar] [CrossRef] [PubMed]
  84. Guerreiro, S.; Toulorge, D.; Hirsch, E.; Marien, M.; Sokoloff, P.; Michel, P.P. Paraxanthine, the primary metabolite of caffeine, provides protection against dopaminergic cell death via stimulation of ryanodine receptor channels. Mol. Pharmacol. 2008, 74, 980–989. [Google Scholar] [CrossRef]
  85. Johnson, R.J.; Kang, D.-H.; Feig, D.; Kivlighn, S.; Kanellis, J.; Watanabe, S.; Tuttle, K.R.; Rodriguez-Iturbe, B.; Herrera-Acosta, J.; Mazzali, M. Is there a pathogenetic role for uric acid in hypertension and cardiovascular and renal disease? Hypertension 2003, 41, 1183–1190. [Google Scholar] [CrossRef] [PubMed]
  86. Rinaldi, P.; Polidori, M.; Metastasio, A.; Mariani, E.; Mattioli, P.; Cherubini, A.; Catani, M.; Cecchetti, R.; Senin, U.; Mecocci, P. Plasma antioxidants are similarly depleted in mild cognitive impairment and in Alzheimer’s disease. Neurobiol. Aging 2003, 24, 915–919. [Google Scholar] [CrossRef]
  87. O’Reilly, E.J.; Gao, X.; Weisskopf, M.G.; Chen, H.; Schwarzschild, M.A.; Spiegelman, D.; Ascherio, A. Plasma urate and Parkinson’s disease in women. Am. J. Epidemiol. 2010, 172, 666–670. [Google Scholar] [CrossRef] [PubMed]
  88. Davis, J.W.; Grandinetti, A.; Waslien, C.I.; Ross, G.W.; White, L.R.; Morens, D.M. Observations on serum uric acid levels and the risk of idiopathic Parkinson’s disease. Am. J. Epidemiol. 1996, 144, 480–484. [Google Scholar] [CrossRef]
  89. Wen, M.; Zhou, B.; Chen, Y.H.; Ma, Z.L.; Gou, Y.; Zhang, C.L.; Yu, W.F.; Jiao, L. Serum uric acid levels in patients with Parkinson’s disease: A meta-analysis. PLoS ONE 2017, 12, e0173731. [Google Scholar] [CrossRef]
  90. Ascherio, A.; LeWitt, P.A.; Xu, K.; Eberly, S.; Watts, A.; Matson, W.R.; Marras, C.; Kieburtz, K.; Rudolph, A.; Bogdanov, M.B.; et al. Urate as a predictor of the rate of clinical decline in Parkinson disease. Arch. Neurol. 2009, 66, 1460–1468. [Google Scholar] [CrossRef]
  91. Weisskopf, M.G.; O’Reilly, E.; Chen, H.; Schwarzschild, M.A.; Ascherio, A. Plasma urate and risk of Parkinson’s disease. Am. J. Epidemiol. 2007, 166, 561–567. [Google Scholar] [CrossRef] [PubMed]
  92. de Lau, L.M.; Koudstaal, P.J.; Hofman, A.; Breteler, M.M. Serum uric acid levels and the risk of Parkinson disease. Ann. Neurol. 2005, 58, 797–800. [Google Scholar] [CrossRef] [PubMed]
  93. Gao, X.; O’Reilly, E.J.; Schwarzschild, M.A.; Ascherio, A. Prospective study of plasma urate and risk of Parkinson disease in men and women. Neurology 2016, 86, 520–526. [Google Scholar] [CrossRef] [PubMed]
  94. Verhaaren, B.F.; Vernooij, M.W.; Dehghan, A.; Vrooman, H.A.; de Boer, R.; Hofman, A.; Witteman, J.C.; Niessen, W.J.; Breteler, M.M.; van der Lugt, A.; et al. The relation of uric acid to brain atrophy and cognition: The Rotterdam Scan Study. Neuroepidemiology 2013, 41, 29–34. [Google Scholar] [CrossRef] [PubMed]
  95. Latourte, A.; Bardin, T.; Richette, P. Uric acid and cognitive decline: A double-edge sword? Curr. Opin. Rheumatol. 2018, 30, 183–187. [Google Scholar] [CrossRef]
  96. Bowman, G.L.; Shannon, J.; Frei, B.; Kaye, J.A.; Quinn, J.F. Uric acid as a CNS antioxidant. J. Alzheimers Dis. 2010, 19, 1331–1336. [Google Scholar] [CrossRef] [PubMed]
  97. Khan, A.A.; Quinn, T.J.; Hewitt, J.; Fan, Y.; Dawson, J. Serum uric acid level and association with cognitive impairment and dementia: Systematic review and meta-analysis. Age 2016, 38, 16. [Google Scholar] [CrossRef]
  98. Hershfield, M.S.; Roberts, L.J., 2nd; Ganson, N.J.; Kelly, S.J.; Santisteban, I.; Scarlett, E.; Jaggers, D.; Sundy, J.S. Treating gout with pegloticase, a PEGylated urate oxidase, provides insight into the importance of uric acid as an antioxidant in vivo. Proc. Natl. Acad. Sci. USA 2010, 107, 14351–14356. [Google Scholar] [CrossRef]
  99. Ames, B.N.; Cathcart, R.; Schwiers, E.; Hochstein, P. Uric acid provides an antioxidant defense in humans against oxidant- and radical-caused aging and cancer: A hypothesis. Proc. Natl. Acad. Sci. USA 1981, 78, 6858–6862. [Google Scholar] [CrossRef]
  100. So, A.; Thorens, B. Uric acid transport and disease. J. Clin. Investig. 2010, 120, 1791–1799. [Google Scholar] [CrossRef]
  101. Huang, T.T.; Hao, D.L.; Wu, B.N.; Mao, L.L.; Zhang, J. Uric acid demonstrates neuroprotective effect on Parkinson’s disease mice through Nrf2-ARE signaling pathway. Biochem. Biophys. Res. Commun. 2017, 493, 1443–1449. [Google Scholar] [CrossRef] [PubMed]
  102. Moriwaki, Y.; Yamamoto, T.; Higashino, K. Enzymes involved in purine metabolism--a review of histochemical localization and functional implications. Histol. Histopathol. 1999, 14, 1321–1340. [Google Scholar] [CrossRef] [PubMed]
  103. Robson, S.C.; Sevigny, J.; Zimmermann, H. The E-NTPDase family of ectonucleotidases: Structure function relationships and pathophysiological significance. Purinergic Signal. 2006, 2, 409–430. [Google Scholar] [CrossRef] [PubMed]
  104. Zimmermann, H. 5’-Nucleotidase: Molecular structure and functional aspects. Biochem. J. 1992, 285 Pt 2, 345–365. [Google Scholar] [CrossRef]
  105. Jarasch, E.D.; Bruder, G.; Heid, H.W. Significance of xanthine oxidase in capillary endothelial cells. Acta Physiol. Scand Suppl. 1986, 548, 39–46. [Google Scholar]
  106. Hille, R.; Nishino, T. Flavoprotein structure and mechanism. 4. Xanthine oxidase and xanthine dehydrogenase. FASEB J. 1995, 9, 995–1003. [Google Scholar] [CrossRef]
  107. Granger, D.N.; Rutili, G.; McCord, J.M. Superoxide radicals in feline intestinal ischemia. Gastroenterology 1981, 81, 22–29. [Google Scholar] [CrossRef]
  108. McCord, J.M.; Roy, R.S. The pathophysiology of superoxide: Roles in inflammation and ischemia. Can. J. Physiol. Pharmacol. 1982, 60, 1346–1352. [Google Scholar] [CrossRef]
  109. Yu, Z.F.; Bruce-Keller, A.J.; Goodman, Y.; Mattson, M.P. Uric acid protects neurons against excitotoxic and metabolic insults in cell culture, and against focal ischemic brain injury in vivo. J. Neurosci. Res. 1998, 53, 613–625. [Google Scholar] [CrossRef]
  110. Amaro, S.; Jimenez-Altayo, F.; Chamorro, A. Uric acid therapy for vasculoprotection in acute ischemic stroke. Brain Circ. 2019, 5, 55–61. [Google Scholar] [CrossRef]
  111. Harris, J.J.; Jolivet, R.; Attwell, D. Synaptic energy use and supply. Neuron 2012, 75, 762–777. [Google Scholar] [CrossRef] [PubMed]
  112. Micheli, V.; Camici, M.; Tozzi, M.G.; Ipata, P.L.; Sestini, S.; Bertelli, M.; Pompucci, G. Neurological disorders of purine and pyrimidine metabolism. Curr. Top Med. Chem. 2011, 11, 923–947. [Google Scholar] [CrossRef] [PubMed]
  113. Belanger, M.; Allaman, I.; Magistretti, P.J. Brain energy metabolism: Focus on astrocyte-neuron metabolic cooperation. Cell Metab. 2011, 14, 724–738. [Google Scholar] [CrossRef] [PubMed]
  114. Dubyak, G.R.; el-Moatassim, C. Signal transduction via P2-purinergic receptors for extracellular ATP and other nucleotides. Am. J. Physiol. 1993, 265 Pt 1, C577–C606. [Google Scholar] [CrossRef]
  115. Khakh, B.S.; North, R.A. P2X receptors as cell-surface ATP sensors in health and disease. Nature 2006, 442, 527–532. [Google Scholar] [CrossRef]
  116. Burnstock, G. Purinergic signalling and disorders of the central nervous system. Nat. Rev. Drug Discov. 2008, 7, 575–590. [Google Scholar] [CrossRef]
  117. Burnstock, G. Historical review: ATP as a neurotransmitter. Trends Pharmacol. Sci. 2006, 27, 166–176. [Google Scholar] [CrossRef]
  118. Neary, J.T.; Rathbone, M.P.; Cattabeni, F.; Abbracchio, M.P.; Burnstock, G. Trophic actions of extracellular nucleotides and nucleosides on glial and neuronal cells. Trends Neurosci. 1996, 19, 13–18. [Google Scholar] [CrossRef]
  119. Melani, A.; Turchi, D.; Vannucchi, M.G.; Cipriani, S.; Gianfriddo, M.; Pedata, F. ATP extracellular concentrations are increased in the rat striatum during in vivo ischemia. Neurochem. Int. 2005, 47, 442–448. [Google Scholar] [CrossRef]
  120. Melani, A.; Corti, F.; Stephan, H.; Muller, C.E.; Donati, C.; Bruni, P.; Vannucchi, M.G.; Pedata, F. Ecto-ATPase inhibition: ATP and adenosine release under physiological and ischemic in vivo conditions in the rat striatum. Exp. Neurol. 2012, 233, 193–204. [Google Scholar] [CrossRef]
  121. Snyder, S.H. Adenosine as a neuromodulator. Annu. Rev. Neurosci. 1985, 8, 103–124. [Google Scholar] [CrossRef] [PubMed]
  122. Ciccarelli, R.; Ballerini, P.; Sabatino, G.; Rathbone, M.P.; D’Onofrio, M.; Caciagli, F.; Di Iorio, P. Involvement of astrocytes in purine-mediated reparative processes in the brain. Int. J. Dev. Neurosci. 2001, 19, 395–414. [Google Scholar] [CrossRef] [PubMed]
  123. Aronsen, L.; Orvoll, E.; Lysaa, R.; Ravna, A.W.; Sager, G. Modulation of high affinity ATP-dependent cyclic nucleotide transporters by specific and non-specific cyclic nucleotide phosphodiesterase inhibitors. Eur. J. Pharmacol. 2014, 745, 249–253. [Google Scholar] [CrossRef] [PubMed]
  124. Ceballos, G.; Tuttle, J.B.; Rubio, R. Differential distribution of purine metabolizing enzymes between glia and neurons. J. Neurochem. 1994, 62, 1144–1153. [Google Scholar] [CrossRef] [PubMed]
  125. Dux, E.; Fastbom, J.; Ungerstedt, U.; Rudolphi, K.; Fredholm, B.B. Protective effect of adenosine and a novel xanthine derivative propentofylline on the cell damage after bilateral carotid occlusion in the gerbil hippocampus. Brain Res. 1990, 516, 248–256. [Google Scholar] [CrossRef] [PubMed]
  126. Braun, N.; Lenz, C.; Gillardon, F.; Zimmermann, M.; Zimmermann, H. Focal cerebral ischemia enhances glial expression of ecto-5’-nucleotidase. Brain Res. 1997, 766, 213–226. [Google Scholar] [CrossRef]
  127. Hasko, G.; Sitkovsky, M.V.; Szabo, C. Immunomodulatory and neuroprotective effects of inosine. Trends Pharmacol. Sci. 2004, 25, 152–157. [Google Scholar] [CrossRef]
  128. Nedeljkovic, N.; Bjelobaba, I.; Subasic, S.; Lavrnja, I.; Pekovic, S.; Stojkov, D.; Vjestica, A.; Rakic, L.; Stojiljkovic, M. Up-regulation of ectonucleotidase activity after cortical stab injury in rats. Cell Biol. Int. 2006, 30, 541–546. [Google Scholar] [CrossRef]
  129. Maples, K.R.; Mason, R.P. Free radical metabolite of uric acid. J. Biol. Chem. 1988, 263, 1709–1712. [Google Scholar] [CrossRef]
  130. Seifar, F.; Dinasarapu, A.R.; Jinnah, H.A. Uric Acid in Parkinson’s Disease: What Is the Connection? Mov. Disord. 2022, 37, 2173–2183. [Google Scholar] [CrossRef]
  131. Davies, K.J.; Sevanian, A.; Muakkassah-Kelly, S.F.; Hochstein, P. Uric acid-iron ion complexes. A new aspect of the antioxidant functions of uric acid. Biochem. J. 1986, 235, 747–754. [Google Scholar] [CrossRef] [PubMed]
  132. Sevanian, A.; Davies, K.J.; Hochstein, P. Serum urate as an antioxidant for ascorbic acid. Am. J. Clin. Nutr. 1991, 54, 1129S–1134S. [Google Scholar] [CrossRef] [PubMed]
  133. Zhu, T.G.; Wang, X.X.; Luo, W.F.; Zhang, Q.L.; Huang, T.T.; Xu, X.S.; Liu, C.F. Protective effects of urate against 6-OHDA-induced cell injury in PC12 cells through antioxidant action. Neurosci. Lett. 2012, 506, 175–179. [Google Scholar] [CrossRef] [PubMed]
  134. Badshah, H.; Ikram, M.; Ali, W.; Ahmad, S.; Hahm, J.R.; Kim, M.O. Caffeine May Abrogate LPS-Induced Oxidative Stress and Neuroinflammation by Regulating Nrf2/TLR4 in Adult Mouse Brains. Biomolecules 2019, 9, 719. [Google Scholar] [CrossRef]
  135. Khan, A.; Ikram, M.; Muhammad, T.; Park, J.; Kim, M.O. Caffeine Modulates Cadmium-Induced Oxidative Stress, Neuroinflammation, and Cognitive Impairments by Regulating Nrf-2/HO-1 In Vivo and In Vitro. J. Clin. Med. 2019, 8, 680. [Google Scholar] [CrossRef]
  136. Quincozes-Santos, A.; Bobermin, L.D.; Souza, D.G.; Bellaver, B.; Goncalves, C.A.; Souza, D.O. Guanosine protects C6 astroglial cells against azide-induced oxidative damage: A putative role of heme oxygenase 1. J. Neurochem. 2014, 130, 61–74. [Google Scholar] [CrossRef]
  137. Souza, D.G.; Bellaver, B.; Bobermin, L.D.; Souza, D.O.; Quincozes-Santos, A. Anti-aging effects of guanosine in glial cells. Purinergic Signal. 2016, 12, 697–706. [Google Scholar] [CrossRef]
  138. Dal-Cim, T.; Poluceno, G.G.; Lanznaster, D.; de Oliveira, K.A.; Nedel, C.B.; Tasca, C.I. Guanosine prevents oxidative damage and glutamate uptake impairment induced by oxygen/glucose deprivation in cortical astrocyte cultures: Involvement of A(1) and A(2A) adenosine receptors and PI3K, MEK, and PKC pathways. Purinergic Signal. 2019, 15, 465–476. [Google Scholar] [CrossRef]
  139. Molz, S.; Dal-Cim, T.; Budni, J.; Martín-de-Saavedra, M.D.; Egea, J.; Romero, A.; del Barrio, L.; Rodrigues, A.L.; López, M.G.; Tasca, C.I. Neuroprotective effect of guanosine against glutamate-induced cell death in rat hippocampal slices is mediated by the phosphatidylinositol-3 kinase/Akt/glycogen synthase kinase 3beta pathway activation and inducible nitric oxide synthase inhibition. J. Neurosci. Res. 2011, 89, 1400–1408. [Google Scholar] [CrossRef]
  140. Miller, L.P.; Jelovich, L.A.; Yao, L.; DaRe, J.; Ugarkar, B.; Foster, A.C. Pre- and peristroke treatment with the adenosine kinase inhibitor, 5’-deoxyiodotubercidin, significantly reduces infarct volume after temporary occlusion of the middle cerebral artery in rats. Neurosci. Lett. 1996, 220, 73–76. [Google Scholar] [CrossRef]
  141. Escartin, C.; Won, S.J.; Malgorn, C.; Auregan, G.; Berman, A.E.; Chen, P.C.; Déglon, N.; Johnson, J.A.; Suh, S.W.; Swanson, R.A. Nuclear factor erythroid 2-related factor 2 facilitates neuronal glutathione synthesis by upregulating neuronal excitatory amino acid transporter 3 expression. J. Neurosci. 2011, 31, 7392–7401. [Google Scholar] [CrossRef] [PubMed]
  142. Shih, A.Y.; Erb, H.; Sun, X.; Toda, S.; Kalivas, P.W.; Murphy, T.H. Cystine/glutamate exchange modulates glutathione supply for neuroprotection from oxidative stress and cell proliferation. J. Neurosci. 2006, 26, 10514–10523. [Google Scholar] [CrossRef] [PubMed]
  143. de Freitas, A.P.; Ferreira, D.D.; Fernandes, A.; Martins, R.S.; Borges-Martins, V.P.; Sathler, M.F.; Dos-Santos-Pereira, M.; Paes-de-Carvalho, R.; Giestal-de-Araujo, E.; de Melo Reis, R.A.; et al. Caffeine alters glutamate-aspartate transporter function and expression in rat retina. Neuroscience 2016, 337, 285–294. [Google Scholar] [CrossRef]
  144. Rosa, P.B.; Bettio, L.E.B.; Neis, V.B.; Moretti, M.; Werle, I.; Leal, R.B.; Rodrigues, A.L.S. The antidepressant-like effect of guanosine is dependent on GSK-3beta inhibition and activation of MAPK/ERK and Nrf2/heme oxygenase-1 signaling pathways. Purinergic Signal. 2019, 15, 491–504. [Google Scholar] [CrossRef]
  145. Zanella, C.A.; Tasca, C.I.; Henley, J.M.; Wilkinson, K.A.; Cimarosti, H.I. Guanosine modulates SUMO2/3-ylation in neurons and astrocytes via adenosine receptors. Purinergic Signal. 2020, 16, 439–450. [Google Scholar] [CrossRef] [PubMed]
  146. Ramos, D.B.; Muller, G.C.; Rocha, G.B.; Dellavia, G.H.; Almeida, R.F.; Pettenuzzo, L.F.; Loureiro, S.O.; Hansel, G.; Horn, Â.C.; Souza, D.O.; et al. Intranasal guanosine administration presents a wide therapeutic time window to reduce brain damage induced by permanent ischemia in rats. Purinergic Signal. 2016, 12, 149–159. [Google Scholar] [CrossRef]
  147. Chen, J.F.; Eltzschig, H.K.; Fredholm, B.B. Adenosine receptors as drug targets--what are the challenges? Nat. Rev. Drug Discov. 2013, 12, 265–286. [Google Scholar] [CrossRef]
  148. Fredholm, B.B.; Abbracchio, M.P.; Burnstock, G.; Daly, J.W.; Harden, T.K.; Jacobson, K.A.; Leff, P.; Williams, M. Nomenclature and classification of purinoceptors. Pharmacol. Rev. 1994, 46, 143–156. [Google Scholar]
  149. Xu, K.; Bastia, E.; Schwarzschild, M. Therapeutic potential of adenosine A(2A) receptor antagonists in Parkinson’s disease. Pharmacol. Ther. 2005, 105, 267–310. [Google Scholar] [CrossRef]
  150. Kostopoulos, G.K.; Phillis, J.W. Purinergic depression of neurons in different areas of the rat brain. Exp. Neurol. 1977, 55 Pt 1, 719–724. [Google Scholar] [CrossRef]
  151. Greene, R.W.; Haas, H.L. The electrophysiology of adenosine in the mammalian central nervous system. Prog. Neurobiol. 1991, 36, 329–341. [Google Scholar] [CrossRef] [PubMed]
  152. von Lubitz, D.K. Adenosine and cerebral ischemia: Therapeutic future or death of a brave concept? Eur. J. Pharmacol. 1999, 371, 85–102. [Google Scholar] [CrossRef] [PubMed]
  153. Rudolphi, K.A.; Schubert, P.; Parkinson, F.E.; Fredholm, B.B. Neuroprotective role of adenosine in cerebral ischaemia. Trends Pharmacol. Sci. 1992, 13, 439–445. [Google Scholar] [CrossRef] [PubMed]
  154. Fredholm, B.B.; Dunwiddie, T.V. How does adenosine inhibit transmitter release? Trends Pharmacol. Sci. 1988, 9, 130–134. [Google Scholar] [CrossRef]
  155. Yawo, H.; Chuhma, N. Preferential inhibition of omega-conotoxin-sensitive presynaptic Ca2+ channels by adenosine autoreceptors. Nature 1993, 365, 256–258. [Google Scholar] [CrossRef]
  156. Wu, L.G.; Saggau, P. Adenosine inhibits evoked synaptic transmission primarily by reducing presynaptic calcium influx in area CA1 of hippocampus. Neuron 1994, 12, 1139–1148. [Google Scholar] [CrossRef]
  157. Segal, M. Intracellular analysis of a postsynaptic action of adenosine in the rat hippocampus. Eur. J. Pharmacol. 1982, 79, 193–199. [Google Scholar] [CrossRef]
  158. Thompson, S.M.; Haas, H.L.; Gahwiler, B.H. Comparison of the actions of adenosine at pre- and postsynaptic receptors in the rat hippocampus in vitro. J. Physiol. 1992, 451, 347–363. [Google Scholar] [CrossRef]
  159. Jacobson, K.A.; Gao, Z.G. Adenosine receptors as therapeutic targets. Nat. Rev. Drug Discov. 2006, 5, 247–264. [Google Scholar] [CrossRef]
  160. Ikeda, K.; Kurokawa, M.; Aoyama, S.; Kuwana, Y. Neuroprotection by adenosine A2A receptor blockade in experimental models of Parkinson’s disease. J. Neurochem. 2002, 80, 262–270. [Google Scholar] [CrossRef]
  161. Kaplan, G.B.; Greenblatt, D.J.; Leduc, B.W.; Thompson, M.L.; Shader, R.I. Relationship of plasma and brain concentrations of caffeine and metabolites to benzodiazepine receptor binding and locomotor activity. J. Pharmacol. Exp. Ther. 1989, 248, 1078–1083. [Google Scholar] [PubMed]
  162. Endesfelder, S.; Weichelt, U.; Strauss, E.; Schlor, A.; Sifringer, M.; Scheuer, T.; Buhrer, C.; Schmitz, T. Neuroprotection by Caffeine in Hyperoxia-Induced Neonatal Brain Injury. Int. J. Mol. Sci. 2017, 18, 187. [Google Scholar] [CrossRef]
  163. Solinas, M.; Ferré, S.; Antoniou, K.; Quarta, D.; Justinova, Z.; Hockemeyer, J.; Pappas, L.A.; Segal, P.N.; Wertheim, C.; Müller, C.E.; et al. Involvement of adenosine A1 receptors in the discriminative-stimulus effects of caffeine in rats. Psychopharmacology 2005, 179, 576–586. [Google Scholar] [CrossRef] [PubMed]
  164. Huang, Z.L.; Qu, W.M.; Eguchi, N.; Chen, J.F.; Schwarzschild, M.A.; Fredholm, B.B.; Urade, Y.; Hayaishi, O. Adenosine A2A, but not A1, receptors mediate the arousal effect of caffeine. Nat. Neurosci. 2005, 8, 858–859. [Google Scholar] [CrossRef] [PubMed]
  165. Horrigan, L.A.; Kelly, J.P.; Connor, T.J. Caffeine suppresses TNF-alpha production via activation of the cyclic AMP/protein kinase A pathway. Int. Immunopharmacol. 2004, 4, 1409–1417. [Google Scholar] [CrossRef] [PubMed]
  166. Rodas, L.; Riera-Sampol, A.; Aguilo, A.; Martinez, S.; Tauler, P. Effects of Habitual Caffeine Intake, Physical Activity Levels, and Sedentary Behavior on the Inflammatory Status in a Healthy Population. Nutrients 2020, 12, 2325. [Google Scholar] [CrossRef]
  167. Xu, K.; Xu, Y.H.; Chen, J.F.; Schwarzschild, M.A. Caffeine’s neuroprotection against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine toxicity shows no tolerance to chronic caffeine administration in mice. Neurosci. Lett. 2002, 322, 13–16. [Google Scholar] [CrossRef]
Figure 1. Regulations of GSH levels in neurons and astrocytes. In both neurons and astrocytes, GSH is synthesized from glutamate (Glu), cysteine (Cys), and glycine (Gly) by two-step enzymatic reactions catalyzed by GCL and GSS. Although Cys is synthesized through the methionine cycle and the transsulfuration pathway, Cys uptake via EAAC1 is the main mechanism supplying Cys to produce intracellular GSH in neurons. In astrocytes, cystine (cysteine disulfide, CysCys) is taken up via xCT and rapidly converted to Cys in intracellular space. In neurons, glutathione disulfide (GSSG), the oxidized form of GSH, is released via MRP1. Astrocytic GSH is released into extracellular space via MRP1. GSSG is recycled to GSH by GSH reductase (GR, Gene ID: 2936), and the released GSH is converted to Cys through the γ-glutamyl cycle. Neurons take up extracellular Cys and use it to synthesize GSH. Other abbreviations: ApN, aminopeptidase N; GLAST, glutamate-aspartate transporter (Gene ID: 6507); Gln, glutamine; GLT-1, glutamate transporter 1 (Gene ID: 6506); GluCys, glutamylcysteine; GluR, glutamate transporter; GPx, glutathione peroxidase (Gene ID: 2876); GS, glutamine synthetase; GST, glutathione S-transferase; CysGly, cysteinylglycine; γ-GT, γ-glutamyl transpeptidase.
Figure 1. Regulations of GSH levels in neurons and astrocytes. In both neurons and astrocytes, GSH is synthesized from glutamate (Glu), cysteine (Cys), and glycine (Gly) by two-step enzymatic reactions catalyzed by GCL and GSS. Although Cys is synthesized through the methionine cycle and the transsulfuration pathway, Cys uptake via EAAC1 is the main mechanism supplying Cys to produce intracellular GSH in neurons. In astrocytes, cystine (cysteine disulfide, CysCys) is taken up via xCT and rapidly converted to Cys in intracellular space. In neurons, glutathione disulfide (GSSG), the oxidized form of GSH, is released via MRP1. Astrocytic GSH is released into extracellular space via MRP1. GSSG is recycled to GSH by GSH reductase (GR, Gene ID: 2936), and the released GSH is converted to Cys through the γ-glutamyl cycle. Neurons take up extracellular Cys and use it to synthesize GSH. Other abbreviations: ApN, aminopeptidase N; GLAST, glutamate-aspartate transporter (Gene ID: 6507); Gln, glutamine; GLT-1, glutamate transporter 1 (Gene ID: 6506); GluCys, glutamylcysteine; GluR, glutamate transporter; GPx, glutathione peroxidase (Gene ID: 2876); GS, glutamine synthetase; GST, glutathione S-transferase; CysGly, cysteinylglycine; γ-GT, γ-glutamyl transpeptidase.
Ijms 24 13067 g001
Figure 2. Purine metabolism in peripheral tissues. (1) ATP is converted to ADP and AMP by nucleoside triphosphate diphospho-hydrolases (NTPDases; CD39), (2) AMP is converted to adenosine by 5′-nucleotidase (5′NT; CD73), (3) adenosine is converted to inosine by adenosine deaminase (ADA), (4) inosine is converted to hypoxanthine by purine nucleoside phosphorylase (PNP), and (5) xanthine oxidase (XO) catalyzes the conversion of hypoxanthine to xanthine and finally to UA. GTP is converted to GDP and GMP by NTPDases. GMP is converted to guanosine by 5′NT, guanosine is transformed into guanine by PNP. Guanine is converted to xanthine by guanine deaminase (GDA, Gene ID: 9615). Caffeine (1,3,7-trimethylxanthine) is demethylated to paraxanthine (PX), theobromine (TB), and theophylline (TP) by cytochrome P450 (CYP). PX, TB, and TP can be further demethylated by CYP to monomethylxanthines such as 1-methylxanthine (1MX), 3MX, and 7MX. The 8-hydroxylation of MXs to form corresponding methyl UAs (MUAs) is mainly catalyzed by XO.
Figure 2. Purine metabolism in peripheral tissues. (1) ATP is converted to ADP and AMP by nucleoside triphosphate diphospho-hydrolases (NTPDases; CD39), (2) AMP is converted to adenosine by 5′-nucleotidase (5′NT; CD73), (3) adenosine is converted to inosine by adenosine deaminase (ADA), (4) inosine is converted to hypoxanthine by purine nucleoside phosphorylase (PNP), and (5) xanthine oxidase (XO) catalyzes the conversion of hypoxanthine to xanthine and finally to UA. GTP is converted to GDP and GMP by NTPDases. GMP is converted to guanosine by 5′NT, guanosine is transformed into guanine by PNP. Guanine is converted to xanthine by guanine deaminase (GDA, Gene ID: 9615). Caffeine (1,3,7-trimethylxanthine) is demethylated to paraxanthine (PX), theobromine (TB), and theophylline (TP) by cytochrome P450 (CYP). PX, TB, and TP can be further demethylated by CYP to monomethylxanthines such as 1-methylxanthine (1MX), 3MX, and 7MX. The 8-hydroxylation of MXs to form corresponding methyl UAs (MUAs) is mainly catalyzed by XO.
Ijms 24 13067 g002
Figure 3. Purine metabolism and neuroprotection. (1) ATP is converted to ADP and AMP by CD39, (2) AMP is converted to adenosine by CD73, (3) adenosine is converted to inosine by ADA. The neuroprotective mechanism of UA is its antioxidant activity, while that of caffeine is A2AAR receptor antagonism. Adenosine reduces neuronal excitability and firing rate via A1AR stimulation. The mechanism of guanosine induced neuroprotection involve A1AR agonism. Upregulation of GSH levels is one of the neuroprotective mechanisms of UA, caffeine, PX, and guanosine. Abbreviations: cAMP, cyclic AMP; MEK, extracellular-signal regulated kinase kinase; P2X, ionotropic purinergic receptor; P2Y, metabotropic purinergic receptor.
Figure 3. Purine metabolism and neuroprotection. (1) ATP is converted to ADP and AMP by CD39, (2) AMP is converted to adenosine by CD73, (3) adenosine is converted to inosine by ADA. The neuroprotective mechanism of UA is its antioxidant activity, while that of caffeine is A2AAR receptor antagonism. Adenosine reduces neuronal excitability and firing rate via A1AR stimulation. The mechanism of guanosine induced neuroprotection involve A1AR agonism. Upregulation of GSH levels is one of the neuroprotective mechanisms of UA, caffeine, PX, and guanosine. Abbreviations: cAMP, cyclic AMP; MEK, extracellular-signal regulated kinase kinase; P2X, ionotropic purinergic receptor; P2Y, metabotropic purinergic receptor.
Ijms 24 13067 g003
Table 1. Neuroprotective mechanisms.
Table 1. Neuroprotective mechanisms.
Mechanism of NeuroprotectionReferences
(i) Suppressing excessive neuronal stimulation
   1. Antagonizing excitatory amino acids[24,25]
   2. Inhibiting neurotransmitter release[26,27]
   3. Promoting neurotransmitter uptake and metabolism[28,29]
(ii) Maintaining antioxidant activity in neurons
   1. GSH syntheses (GCL and GSS)[30,31,32,33]
   2. Cysteine uptake (EAAC1)[34,35,36]
   3. GSH supply from astrocyte (GCL, xCT, and MRP1)[28,37,38,39]
   4. Other antioxidants (ascorbate, UA, and α-tocopherol)[11,40,41,42,43]
(iii) Detoxifying xenobiotics
   1. Induction of phase I enzymes[44,45]
   2. Induction of phase II enzymes[46]
   3. Exclusion via MRPs[47,48]
Abbreviations: EAAC1, excitatory amino acid carrier protein 1; GCL, γ-glutamyl cysteine ligase; GSH, glutathione; GSS, GSH synthetase; MRP1, multidrug resistance protein 1; MRPs, multidrug resistance proteins; UA, uric acid; xCT, cystine/glutamate antiporter.
Table 2. Affinity of purine derivatives at the targets.
Table 2. Affinity of purine derivatives at the targets.
Purine DerivativesAdenosine Receptor SubtypesReferences
A1ARA2AARA2BARA3AR
Antagonist potency for AR (Kb, µM)[62]
Caffeine33.812.315.5>100
Paraxanthine15.85.35.5>100
Theophylline8.97.94.8>100
Agonist potency for AR (EC50, µM)[62]
Adenosine0.310.7323.50.29
Inosine290inactiveinactive0.25
Antagonist potency for AR (Ki, µM) in striatum[65]
Caffeine20.58.6
Paraxanthine5.07.5
Theophylline4.79.8
Theobromine96.5109.5
Antagonist potency for AR (Ki, µM)[66]
Theophylline6.81.77.986
Agonist potency for AR * (IC50, µM)[67]
Adenosine0.0700.15015.46.5
Inhibition of AR (IC50, µM)[68]
UA >5300
Caffeine107
Other TargetsReferences
PDERyR
Inhibition of AR (IC50, µM)[69]
Caffeine400
Caffeine 2000
* Affinity for AR was analyzed by adenylate cyclase inhibition. Abbreviations: A1AR, adenosine A1 receptor; A2AAR, adenosine A2A receptor; A2BAR, adenosine A2B receptor; A3AR, adenosine A3 receptor, PDE, phosphodiesterase; RyR, ryanodine receptor.
Table 3. Neuroprotective effects of purine derivatives via regulating GSH synthesis and AR.
Table 3. Neuroprotective effects of purine derivatives via regulating GSH synthesis and AR.
Treatment or StimulusModelAnalysis (Neuroprotective Activity)Signal TransductionReferences
UA6-OHDA-treated PC12 cells (UA 200–400 µM)LDH release↓, MDA↓, and 8-OHdG levels↓SOD↑ and GSH↑[133]
UA6-OHDA-induced striatum lesioned rat (twice daily UA 200 mg/kg i.p. 10 days)DA neuron loss↓, behavioral deficit↓, and MDA↓SOD↑ and GSH↑[8]
UA6-OHDA-treated SH-SY5Y cells (200 µM UA)cell viability↑PI3K↑ and Akt/GSK3β↑[8]
UAMPTP treated Parkinson’s disease model mouse (UA 250 mg/kg i.p.)Recovery of behavioral and cognitive function. DA neuron loss↓, GFAP+ astrocyte↓, and MDA↓Nrf2 (nuc translocation)↑ GCLC/NQO1/HO-1↑, SOD↑, CAT↑, and GSH↑[101]
UACerebral ischemia/reperfusion model rat (UA 16 mg/kg i.v.)TUNEL+ cell↓, MDA↓, carbonyl groups↓, and 8-OHdG levels↓Nrf2/BDNF and NGF levels↑[9]
UAMouse cortical astrocyte culture (100 µM UA)Oxidative stress-induced DA neuronal cell death↓Nuclear translocation of Nrf2↑, GCLM/NQO1/HO-1↑, and GSH synthesis and release↑[37]
CaffeineLPS-induced oxidative stress model mouse (Caffeine 30 mg/kg/day i.p. 4 weeks)Apoptotic cell death↓ and synaptic dysfunction↓Nrf2/HO-1↑ and TLR4/p-NF-κB/p-JNK↓[134]
CaffeineCadmium-induced cognitive deficits model mouse (caffeine 30 mg/kg i.p. 2 weeks)Neuronal cell loss↓ and synaptic dysfunction↓Nrf2/HO-1↑[135]
CaffeineHT-22 and BV-2 cells (100 µM caffeine)ROS↓ and lipid peroxidation↓Nrf2/HO-1↑[135]
Caffeine and UASIN-1 treated mouse (caffeine 10 mg/kg i.p. and UA10 mg/kg i.p.).Nitrotyrosine levels in hippocampal slice↓Cys uptake via EAAC1↑ GSH↑[20]
ParaxanthineH2O2 treated SH-SY5Y cells (10–100 µM paraxantine)LDH release↓Cys uptake via EAAC1↑ GSH↑[21]
GuanosineC6 astroglia and adult rat astrocyte culture (100 µM guanosine)Azide-induced cytotoxicity↓HO-1↑, GS/GR/GCL↑ and GSH↑[136,137]
GuanosineCortical astrocyte culture (10 µM guanosine)Oxygen/glucose deprivation and reoxygenation-induced cell death↓A1AR agonism and A2AR antagonism, PI3K/Akt↑, and MEK/ERK↑[138]
GuanosineRat hippocampal slices (100 µM guanosine)Glu-induced cell death↓PI3K/Akt//GSK3↑[139]
AdenosineCerebral ischemia/reperfusion model rat (treatment with adenosine kinase inhibitor)Infarct volume↓A1AR agonism[2,140]
CaffeineMPTP treated mice (Caffeine 10–20 mg/kg, i.p.)Striatal DA depletion↓A2AAR antagonism[5]
Caffeine, paraxanthine, and theophyllineMPTP treated mice (caffeine 10 mg/kg, Paraxanthine 10–30 mg/kg, Theophylline 10–20 mg/kg, i.p.)Striatal DA depletion↓ [6]
TheobromineFat-enriched diet-induced cognitive deficits model rat (Theobromine 30 mg/L in drinking water)Improve cognitive functions and Aβ and IL-1β levels in brain↓A1R mRNA and protein level↑[7]
ParaxanthineMPP+ treated rat DA neuron culture (800 µM paraxanthine)DA neuron loss↓Ryanodine receptor activation[84]
Abbreviations: Akt, protein kinase B; Aβ, amyloid-beta protein; BDNF, brain-derived neurotrophic factor; CAT, catalase; DA, dopamine; ERK, extracellular-signal regulated kinase; GCLC, GCL catalytic subunit; GCLM, GCL modifier subunit; GSK3, glycogen synthase kinase 3; HO-1, hemeoxygenase-1; IL-1β, interleukin-1β; i.v., intravenously injection; LDH, lactate dehydrogenase; LPS, lipopolysaccharide; MDA, Malondialdehyde; MPP+, 1-methyl-4-phenylpyridinium; MPTP, 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine, NGF, nerve growth factor; Nrf2, nuclear factor erythroid-2-related factor 2; NQO1, NAD(P)H quinone oxidoreductase 1; p-JNK, phosphor-c-Jun n-terminal kinase; PI3K, Phosphoinositide 3-Kinase; p-NF-κB, phosphor-NF-κB; ROS, reactive oxygen species; SIN-1, 3-morpholinosydnonimine; SOD, superoxide dismutase; TLR4, toll-like receptor 4,; TUNEL, Terminal deoxynucleotidyl transferase-mediated dNTP nick end labeling; 6-OHDA, 6-hydroxydopamine; 8-OHdG, 8-hydroxyl-2′-deoxyguanosine. Cell lines: BV-2, murine microglia cell line; HT22 cells, mouse embryonic fibroblasts; PC12 cells, rat pheochromocytoma cell line; SH-SY5Y cells, human neuroblastoma. ↑ and ↓ indicate an increase and decrease of the object, respectively.
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

Matsumura, N.; Aoyama, K. Glutathione-Mediated Neuroprotective Effect of Purine Derivatives. Int. J. Mol. Sci. 2023, 24, 13067. https://doi.org/10.3390/ijms241713067

AMA Style

Matsumura N, Aoyama K. Glutathione-Mediated Neuroprotective Effect of Purine Derivatives. International Journal of Molecular Sciences. 2023; 24(17):13067. https://doi.org/10.3390/ijms241713067

Chicago/Turabian Style

Matsumura, Nobuko, and Koji Aoyama. 2023. "Glutathione-Mediated Neuroprotective Effect of Purine Derivatives" International Journal of Molecular Sciences 24, no. 17: 13067. https://doi.org/10.3390/ijms241713067

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