Next Article in Journal
Toward Ameliorating Insulin Resistance: Targeting a Novel PAK1 Signaling Pathway Required for Skeletal Muscle Mitochondrial Function
Next Article in Special Issue
Effects of Redox Homeostasis and Mitochondrial Damage on Alzheimer’s Disease
Previous Article in Journal
Flt3 Activation Mitigates Mitochondrial Fragmentation and Heart Dysfunction through Rebalanced L-OPA1 Processing by Hindering the Interaction between Acetylated p53 and PHB2 in Cardiac Remodeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Lactate: A Theranostic Biomarker for Metabolic Psychiatry?

1
Clinical Imaging Sciences Centre, Brighton and Sussex Medical School, University of Sussex, Falmer BN1 9RR, UK
2
Department of Clinical Neuroscience, Brighton and Sussex Medical School, University of Sussex, Falmer BN1 9RR, UK
3
Independent Researcher, Florianópolis 88062-300, Brazil
4
School of Life Sciences, University of Sussex, Falmer BN1 9RR, UK
*
Authors to whom correspondence should be addressed.
Antioxidants 2023, 12(9), 1656; https://doi.org/10.3390/antiox12091656
Submission received: 5 July 2023 / Revised: 1 August 2023 / Accepted: 16 August 2023 / Published: 22 August 2023

Abstract

:
Alterations in neurometabolism and mitochondria are implicated in the pathophysiology of psychiatric conditions such as mood disorders and schizophrenia. Thus, developing objective biomarkers related to brain mitochondrial function is crucial for the development of interventions, such as central nervous system penetrating agents that target brain health. Lactate, a major circulatory fuel source that can be produced and utilized by the brain and body, is presented as a theranostic biomarker for neurometabolic dysfunction in psychiatric conditions. This concept is based on three key properties of lactate that make it an intriguing metabolic intermediate with implications for this field: Firstly, the lactate response to various stimuli, including physiological or psychological stress, represents a quantifiable and dynamic marker that reflects metabolic and mitochondrial health. Second, lactate concentration in the brain is tightly regulated according to the sleep–wake cycle, the dysregulation of which is implicated in both metabolic and mood disorders. Third, lactate universally integrates arousal behaviours, pH, cellular metabolism, redox states, oxidative stress, and inflammation, and can signal and encode this information via intra- and extracellular pathways in the brain. In this review, we expand on the above properties of lactate and discuss the methodological developments and rationale for the use of functional magnetic resonance spectroscopy for in vivo monitoring of brain lactate. We conclude that accurate and dynamic assessment of brain lactate responses might contribute to the development of novel and personalized therapies that improve mitochondrial health in psychiatric disorders and other conditions associated with neurometabolic dysfunction.

1. Introduction

Metabolic psychiatry is an emerging field which aims to improve outcomes amongst those suffering with psychiatric disorders by therapeutically targeting metabolism. This subspecialty is grounded in the bidirectional association of metabolic (e.g., obesity and type 2 diabetes) and mood and affective disorders, such as bipolar disorder (BPD), major depressive disorder (MDD), and schizophrenia (SCZ) [1].
Cellular metabolic homeostasis is crucially regulated by mitochondria, bi-genomic organelles which integrate environmental signals and coordinate cellular fate, circadian rhythms, inflammation, and immunity [2]. The role of mitochondria in human health and disease is increasingly recognised [3], and it is particularly relevant for understanding the pathophysiology of brain disorders. In the central nervous system (CNS), their function is universal and complex due to the exceptional energetic demands that is required by higher-order brain activity [4]. In rare disorders caused by mutations in genes that encode for mitochondrial proteins, neurological impairment is common, and there is an elevated risk of developing comorbid psychiatric disorders [5]. The complexity is amplified by the fact that mitochondria accumulate and express mitochondrial (mt)DNA mutations throughout their lifespan, are involved in brain ageing and neurogenerative processes [6], and respond to psychological stress [7].

1.1. The Role of Mitochondria in the Pathogenesis of Psychiatric Disorders

The role of mitochondria and neurometabolism in the aetiopathophysiology of mood and affective disorders has been reviewed extensively [8,9,10]. Pre-clinical, post-mortem, cerebrospinal fluid (CSF) sampling, and imaging studies have yielded valuable insight in MDD, BPD, and SCZ thus far. Of note, chronic mild stress dissipates membrane potential in mitochondria and damages brain mitochondrial ultrastructure in mice [11]. In a pre-natal stress mouse model of depression, the rate of glycolysis is increased, and oxidative metabolism is reduced [12]. This is consistent with the finding that in depressed adults, there are elevated concentrations of the mitochondrial fuel lactate in the anterior cingulate cortex (ACC) [13] and CSF [14]. In fact, the finding of elevated CSF lactate is consistent across MDD, BPD [15], and SCZ [14]. Despite the nomenclature, the gene Disrupted In Schizophrenia 1 (DISC1) confers risk for all mental illnesses [16]. DISC1 is implicated in astrocytic energy metabolism and regulates glucose uptake and lactate production [17]. In SCZ, there are abnormalities of both glycolytic and oxidative enzyme function [18] and a reduction in axonal mitochondrial density in the ACC [19], which is consistent with in vivo imaging showing a reduced creatine kinase (CK) reaction rate for ATP synthesis and pH [20]. Elevated ACC lactate levels negatively correlate with both general cognition and functional capacity in SCZ [21]. In patients with first episode psychosis, peripheral blood mononuclear cells (PBMCs) have significantly increased protein expression of both lactate dehydrogenase B (LDHB) and glucose-6-phosphate isomerase (GPI), highlighting the need to consider mitochondria and metabolic markers in the brain and the body [22].
In BPD, mitochondrial dysfunction may play a particularly important role. The post-mortem profiling of BPD patients has revealed abnormal mitochondrial morphology [23] and impaired prefrontal cortex (PFC) electron transport chain (ETC) complex assembly, specifically mitochondrial complex 1 (MC1), with increased oxidative damage [24,25]. There is also reduced hippocampal expression of the MC1 subunit gene, NDUFV2 [26]. Changes in mtDNA across several brain regions have been mapped in BPD, with data indicating the vulnerability of different brain regions to energetic demands, for example, with increased and decreased mtDNA levels present in the hippocampus and PFC compared to control post-mortem brains, respectively [27]. Phosphorus magnetic resonance spectroscopic (MRS) studies in BPD patients identified alterations in phosphocreatine [28], creatine kinase reaction rate constant [29], and ATP levels after stimulation [30], whilst proton MRS imaging studies have consistently found elevated brain lactate in BPD (discussed further in the following sections), highlighting the fact that neurometabolic alterations can be detected post-mortem as well as in the living brain of BPD patients [31,32].
Whilst much research has focused on differences between people with these distinct diagnoses and non-affected participants, it would also be important to consider the broader impact of mitochondrial dysfunction on human behaviour and health and transcend the traditional categorical diagnostic criteria and boundaries for psychiatric and neurological diseases. Ideally, mitochondrial dysfunction should be conceptualised and studied through the framework of the six domains of the Research Domain Criteria (RDoC) [33] (negative valence, positive valence, cognitive, social, arousal, and sensorimotor systems) [34]. Either way, there is a clear and unmet need in psychiatry for biomarkers that can be mapped across these domains and linked back to quantitative measures of mitochondrial function.

1.2. Potential of Cerebral Lactate as a Neurometabolic Biomarker

Lactate, a major circulatory fuel source that can be produced and utilized by the brain and body, is an excellent candidate to serve as a theranostic biomarker for neurometabolic dysfunction. In sports and exercise medicine, the ability to clear fatty acids and lactate from blood over time is a proxy assessment of systemic oxidative capacity [35], i.e., the efficiency of mitochondrial function at the level of the whole body [36]. Systemic lactate measurements have also been found to inform mechanisms of psychiatric phenomena: early work studying blood lactate dynamics observed that stress precipitated ‘unusual responses at lower stimulus levels’ in those suffering from ‘neurocirculatory asthenia’ (a historical diagnostic category now superseded by elements of anxiety and panic disorder) compared to control participants [37]. Exercising anxious subjects exhibited abnormalities of lactate metabolism that were comparable to those with cardiovascular disease [38,39,40]. This prompted the ‘lactate hypothesis of anxiety’ and the characterisation of exogenous sodium lactate as a panicogenic agent [41]. More recently, lactate has also been purported to have antidepressant properties [42], and the field has received much attention with lactate going from ‘from rags to riches’ [43], once ‘the ugly duckling of metabolism’ [44], now ‘a phoenix risen’ [45].
Despite its potential as a biomarker, the peripheral concentration of lactate lacks the ability to adequately reflect mitochondrial function within the brain (and the body), due to the heterogenous distribution of mitochondrial populations between cells, tissues, and organs with different heteroplasmy rates [46]. In fact, it is the tissue specific heteroplasmy rate that may dictate which systems are implicated in mitochondrial disease and influence the profile of neuropsychological symptoms [47]. Despite decades of development and progress in quantitative neurometabolic imaging techniques including positron emission tomography (PET), or magnetic resonance spectroscopy (MRS)-derived measurements of cerebral glucose or oxygen metabolic rate, there is a case to be made for integrating these with tools that have greater specificity for the pathological processes of interest. The MRS techniques enable spatiotemporal resolution of metabolites relevant to bioenergetics, and the characterisation of brain lactate dynamic responses through functional MRS (fMRS) is a promising tool to enable the direct assessment of mitochondrial homeostatic responses to neurometabolic stress [48]. Although measurement of lactate with MRS is challenging due to a low signal-to-noise ratio (SNR) in vivo, we opine that lactate has a key role that enables an integrated and precise approach for appreciating and understanding energy allocation and mitochondrial efficiency within the CNS that goes beyond oxygen and blood flow measurement [49]. Indeed, changes in brain lactate metabolism measured with MRS have already been used to characterise and monitor the response to pharmacological therapy in psychiatric conditions (see Section 4).

1.3. Aim of the Review

In the present review, we hope to drive attention to the potential and under-explored significance of brain lactate as a dynamic biomarker for neurometabolic dysfunction that is relevant to many psychiatric conditions and the RDoC neurobehavioural domains. We will define characteristics of lactate in cells and the brain, outlining the molecular and physiological mechanisms by which it may signal and modulate single neurons, brain nuclei, cognition, and behaviour. We outline how the spatiotemporal regulation of lactate dynamic responses may reflect the state of arousal in the nervous system and consider the implications of these altered lactate dynamics that may occur due to mitochondrial and metabolic dysfunction.
These observations highlight the possible clinical utility and characterisation of brain lactate metabolism as a potential therapeutic target, which could inform future empirical testing of this hypothesis. If confirmed via rigorous experimental studies through testing the effect of treatment interventions, the characterisation of brain lactate dynamic responses with fMRS could represent an in vivo precision medicine tool to track the effects of novel or pre-existing therapies that target brain metabolism and mitochondrial function, including pharmacological agents but also personalized exercise protocols, dietary interventions, and nutraceuticals.

2. Lactate Neurophysiology

Lactate is a three-carbon monocarboxylate that exists predominantly as a negatively charged anion at a physiological pH [50]. It is chiral and can exist as three isomers: DL-lactate, D-lactate, and L-lactate. In the brain, L-lactate (referred to as lactate from now on, unless otherwise specified) is the dominant form [51]. It is produced intracellularly with NAD+ via glycolysis and exists in equilibrium with the redox couple pyruvate and NADH [44]. Long thought of as a dead-end waste product produced in response to hypoxia, it is now considered to be the obligatory end-product of glycolysis and a crucial substrate that fuels mitochondrial bioenergetics [43]. Lactate is enriched around mitochondria with concentrations in the range of ~268–1025 µM (<100 µM in the nucleus/cytosol) [52]. Here, it acts as an intermediary between glycolysis and oxidative metabolism, generating adenosine triphosphate (ATP) via pyruvate, or shuttling between other cells, tissues, and organs to fuel the tricarboxylic acid cycle (TCA) or replenish glucose [53]. Lactate is the most abundant carbon metabolite in the blood after glucose, with an average circulating concentration of 1 mM at rest [54,55]. When the rate of systemic uptake is limited by impaired oxidative metabolism, lactate concentrations become elevated, as evidenced by primary mitochondrial disease [56]. Similar observations can be made in obesity and in those with metabolic syndrome (MetS) [57]. Conversely, when the mitochondrial capacity of muscle and other tissues is enhanced by physical training, the clearance of lactate can become more efficient [58,59]. With these two ends of the spectrum in mind, Brooks and San-Millán developed a clinical tool and proxy for measuring systemic mitochondrial health using serial lactate measurements under incremental exercise intensity testing [35].
The concentration of lactate in the brain exhibits a robust circadian rhythm [60]. In rodent models, the baseline concentration of 350–420 μM measured with biosensors cycled between −16.2% and +53% [61]. This rhythm was found to be actively regulated by the drainage systems that can clear lactate during natural sleep [62]. Microdialysis data show that cortical lactate is lower during sleep, and lower still during deep sleep [63,64]. These state-dependent changes are controlled both by interstitial clearance mechanisms and an activity-dependent reduction in glycolysis, which can be induced with anaesthesia [65]. To reinforce this relationship, peripheral administration of lactate can trigger wakening [66], and when mice are startled from sleep, lactate rises as much as 50% above baseline [61]. Brain lactate concentration is highly correlated with state-dependent changes in arousal and the concentration of norepinephrine (NE), and it oscillates across the sleep–wake cycle, which is altered by circadian disruption and aging [60,67]. Although glycolytic rates are lower during sleep than during wakefulness [68], the brain remains a net exporter of lactate, at least in healthy individuals at rest [69].
Neurons and astrocytes both carry out glycolysis and export lactate [70]. Astrocytes contain stores of glycogen and exhibit a more glycolytic phenotype and hence produce the majority of cerebral lactate [71,72,73]. Somatosensory stimulation triggers the activation of neurons and increases the local concentration of lactate in the brain, linking transmission and neurometabolism [74,75]. Still somewhat debated, astrocytes provide metabolic support for working neurons, a concept underpinned by the astrocytic-neuronal lactate shuttle (ANLS) [76,77]. Of particular relevance to psychiatry with the RDoC domains in mind, data implicate the ANLS in arousal [78], stress responses [79], memory formation [80], and memory processing [81]. Similarly, chronic stress depletes astrocytic glycogen, leading to impaired synaptic plasticity [82] and depression-like behaviour [83].

3. Lactate Signalling in the Brain

When the systemic lactate concentration is elevated from baseline such as during strenuous physical activity the brain becomes a net consumer (above ~4 mM) [84,85,86]. The fate of this excess lactate within the brain is unclear, yet virtually every function of the brain can be influenced by it. It is important to highlight that depending on the metabolic profile of the individual, differential amounts of lactate may be produced and enter the brain over a time course which may have signalling implications [35]. Lactate displays mechanisms of action both extra- and intracellularly that are dependent on the intensity of the stimulus or level of activity, the balance between production and clearance by glycolysis and mitochondrial efficiency, the concentration gradient and diffusion properties of the tissue [87]. The molecular and intracellular mechanisms, tissue level effects, implicated groups of neurons, and RDoC-mapped neurobehavioural effects of lactate in the brain are described in the section below and summarised in Figure 1.
Extracellular lactate in the brain is associated with a relative reduction in pH [89]. Monocarboxylate transporters (MCTs), specifically MCT1-4, co-transport lactate with a proton, directly coupling lactate concentration and pH across compartments [90]. This intrinsic property of MCT1 on the blood–brain barrier therefore dictates that any lactate entering the cerebral compartment results in a relative reduction in pH. Generally, pH regulates synaptic transmission and neuronal firing via calcium voltage-gated channels and glutamate receptors [91]. Acid-sensing ion channels (ASICs) sense changes in pH and are distributed throughout the brain [92]. ASICs in the amygdala have been proposed as the target for lactate-induced panic [93]. Three other important brain areas also sense pH changes. The first is hypothalamic orexin neurons, a key energy sensing hub of the brain [94]. These neurons mediate autonomic arousal and influence wakefulness with widespread projections to major arousal-related centres including the thalamic nucleus, central grey, raphe nuclei, and locus coeruleus (LC) [95]. The second and third are serotonergic raphe nuclei [96] and the noradrenergic LC [97]. These nuclei have extensive projections that regulate arousal and neurovascular tone in the brain and body, and it almost goes without saying that both NE and serotonin are implicated broadly in mood and affective disorders.
Extracellular lactate can signal via membrane-bound receptors in the brain. In the brain, the hydroxycarboxylic acid receptor 1 (HCAR1) is found on plasma membranes of excitatory synapses within the hippocampus, cerebellum, and cortex [98]. It is also widely expressed by interneurons [99]. The activation of HCAR1 by lactate may decrease neuronal excitability and firing frequency [100], consistent with other data suggesting that lactate may act to suppress neuronal firing [101]. On the other hand, lactate appears to increase motor cortex excitability [102], and it may modulate high-frequency gamma oscillations [103]. These differences are likely related to temporal changes across distinct anatomical regions, with one group suggesting that activation of HCAR1 may aid memory consolidation after event-related activity [104]. Indeed, lactate can promote neurogenesis and angiogenesis via HCAR1 [105,106,107]. As of yet, no data link lactate receptors directly to mood and affective disorders, but HCAR1 does exert downstream effects on cyclic adenosine monophosphate (cAMP), and excessively elevated levels of cAMP are implicated in schizophrenia, fatigue, stress, cognitive impairment, and Alzheimer’s disease [99]. Other receptors for lactate are likely to exist, but so far remain elusive [108]. Perhaps most importantly, lactate directly stimulates LC neurons and can trigger an increase systemic blood pressure [109]. Even oral ingestion of lactate triggers the expression of tyrosine hydroxylase within LC neurons, the enzyme necessary for catecholamine synthesis [110]. Thus, extracellular lactate may have profound effects on neurocortical activity and arousal-related behaviours relevant to psychiatry.
Brain lactate has important actions intracellularly where it maintains the redox state with pyruvate and the redox couple NAD+/NADH. In the brain, blood flow is modulated by this redox couple acting via a prostaglandin transporter that can respond to regional fluctuations in lactate [111] and enable efficient delivery of oxygen to areas that are most active [112]. Ca2+ is a key ion that can modulate mitochondrial metabolism in astrocytes [113], and similarly responds to the lactate-coupled NAD+/NADH ratio, modulating the astrocytic response to dopamine [114].
As a fuel of oxidative metabolism that can shuttle between mitochondria throughout the body, it would follow that a lactate sensor would exist in the brain to inform global energy balance and coupled behaviours. Indeed, the orexin neurons of the hypothalamus appear to act as the central lactate sensor, coupling the local concentration of lactate with their activity in an ATP-dependent manner [115].
Finally, great interest has been generated by the discovery of histone lactylation, a novel form of epigenetic modification [116]. Both extra- and intracellularly derived lactate driven by glycolysis or impaired mitochondrial function result in lysine lactylation and influence gene expression [117], leading to metabolic reprogramming [118]. One group found that neural excitation stimulated by social defeat stress in rodents precipitated H1 histone lactylation that, fascinatingly, correlated with a decrease in social behaviours [119]. Overall, these data highlight that lactate is a significant signalling molecule in the brain with pleiotropic effects relevant to the field of psychiatry.

4. Brain Lactate Dynamics as a Putative Therapeutic Biomarker

Chronic elevation of lactate raises ROS leading to mitochondrial dysfunction in a harmful feedback loop [120]. Thus, improving mitochondrial function and the associated lactate dynamics may potentially be useful for monitoring the metabolic effects of clinical therapy over the longer term. Paradigmatic examples of similar applications supporting this hypothesis are in the progressive neurodegenerative disorder Huntington’s disease, where elevated brain lactate could be reduced by Coenzyme Q10 (CoQ10) [121]. Similarly, in primary mitochondrial disorders where elevated lactate can be diagnostic, therapeutic approaches which target mitochondrial metabolism including CoQ10, riboflavin, and other B vitamins are trialled in order to improve mitochondrial function [122]. The notion that lactate might be a useful therapeutic biomarker in clinically defined populations such as those with metabolic syndrome is supported by evidence that measurements of peripheral lactate dynamics are sensitive to detect the effects of therapy associated with improvements in mitochondrial oxidative capacity, such as the reversal of a pre-type 2 diabetes through personalised exercise prescription [3,35]. The observations that systemic lactate is a proxy for mitochondrial efficiency that is sensitive to treatment effects, together with well documented evidence of mitochondrial brain abnormalities implicated in the aetiopathogenesis of psychiatric disorders, support our hypothesis that the ability to measure dynamic changes in lactate in the brain is likely to be a valuable avenue of research for psychiatric treatment. Data from interventional studies in psychiatric conditions, indicating that effective psychiatric treatments lead to brain lactate alterations, further reinforce this idea: in BPD patients, six weeks of lithium therapy was associated with a reduction in the concentration of lactate measured in the ACC to levels seen in healthy controls [123]. Twelve weeks of treatment with the anti-psychotic quetiapine correlated with a reduction in lactate in frontal areas of the brain, and responders exhibited greater decreases in brain lactate compared to non-responders [124]. Venlafaxine, a commonly prescribed antidepressant drug, was recently shown to have the ability to rescue metabolic changes in the hippocampus of rats exposed to chronic mild stress, raising the relative abundance of both glucose and lactate in cells compared to vehicle control [125]. In a small cohort of geriatric patients with BPD, just four weeks of CoQ10 supplementation improved depression scores measured with the Montgomery–Asberg Depression Rating [126]. With the knowledge that many psychiatric medications, in particular anti-psychotics, can impair or worsen metabolic health in patients, it is vital that old and new therapeutic approaches alike consider the metabolic effects.

Therapeutic Interventions Targeting Brain Mitochondria and Lactate Dynamics

In this section, we will discuss pathways and targets that could enable manipulation of lactate as a by-product of improving metabolism, most notably through improvements in mitochondrial oxidative efficiency. Many factors influence mitochondria, all of which present themselves as targets that could be leveraged for therapeutic benefit (Figure 2).
  • Physical exercise
It is well documented that physical exercise can have a positive impact on psychiatric symptoms. The physiological significance of lactate derived from exercise in the brain and the body is an active topic of discussion [127]. Just a single bout of high intensity exercise in mice corresponding to a blood lactate concentration of greater than 12 mM increased mtDNA copy number and the expression of Peroxisome proliferator-activated receptor Gamma Coactivator 1-alpha (PGC-1α) (the master regulator of mitochondrial biogenesis) in the hippocampus [128]. The inhibition of either MCT2 or the histone deacetylase (HDAC) Silent Information Regulator 1 (SIRT1) prevented this increase in hippocampal PGC-1α protein level, indicating that the observed effects are dependent on lactate transport acting via SIRT1 [129]. In contrast, low- or moderate-intensity exercise (3–4 mM) did not induce these changes in a single bout [128], consistent with data that repeated bouts are necessary at lower or moderate intensities [130]. In humans, low and moderate training intensity has been applied to specifically target mitochondrial function, improving plasma lactate and fatty acid clearance over the course of an individualised 12 month programme [3]. Strikingly, this approach has already been trialled as an adjuvant cancer therapy in order to target altered lactate metabolism, a metabolic hallmark of cancer [131]. With lactate metabolism in mind, the clinician has a new tool for tailoring the exercise prescription to alter and improve lactate dynamics on a case-by-case basis.
  • Pharmacological enhancement of mitochondrial function
The health of mitochondria in the brain is dependent on the processes that control their dynamics, number, and quality: mitophagy, biogenesis, fission, fusion, as well as the rate of mtDNA mutations [132]. Many compounds are under investigation for their beneficial effects on diverse clinical conditions and can target pathways that regulate mitochondria. Major large-scale randomised, placebo-controlled trials for neuropsychiatric conditions have already been called for by other groups due to the links between nitroxidative stress and brain bioenergetic failure, and compounds which have strong theoretical application with some convincing empirical data should be investigated further [133]. Key molecular nodes include SIRT1 and PGA-1α, which have been discussed above. Other important proteins that regulate mitochondria include the Peroxisome Proliferator-Activated Receptor (PPAR) transcription factors shown to regulate mitochondrial function in the brain [134]. Another is the conserved AMP-activated Protein Kinase (AMPK), which senses cellular energy status and can modulate several key elements of mitochondrial dynamics and quality [135]. Nuclear Factor erythroid 2-related Factor 2 (NRF2) is another transcription factor that coordinates the cellular antioxidant response but has been shown to broadly regulate mitochondrial quality control [136]. Several other less-specific targets are listed alongside a list of the most promising compounds known to cross the blood–brain barrier in Table 1. As a fundamental component of the electron transport chain, it is not surprising that Coenzyme Q10 has been shown to lower lactate levels in the brain in patients with neurological disorders [137]. Resveratrol, a polyphenol found in red grapes, significantly reduces the cerebral lactate accumulation in response to ischaemia in a mouse model of ischaemic brain injury [138]. Interestingly, metformin appears to have beneficial effects on a number of metabolic parameters but is documented to be an inhibitor of mitochondrial complex I and associated with lactic acidosis [139]. Overall, we feel that in measuring the dynamic lactate response to these pharmaceutical agents, and doing so over time, will yield valuable information that enables precise and personalised approaches to metabolic psychiatry.
  • Nutritional and dietary intervention
Alongside the micronutrients outlined in Table 1, dietary practices that include intermittent fasting (IF), calorie restriction (CR), and the ketogenic diet (KD) are purported to have an impact on mitochondrial function and quality control. Both IF and CR alter mitochondrial dynamics and improve biogenesis and ROS production [221]. Long-term CR in rats reduces the accumulation of liver mtDNA oxidative damage as a result of altered ROS production [222]. Fasting-induced ketosis acutely increases brain lactate [223], which does not occur with infusion of exogenous β-hydroxybutyrate [224], thus being a phenomenon speculated to be underpinned by the chronic adaptive upregulation of brain MCTs. Indeed, the anti-convulsant effects of the KD and induced mitochondrial biogenesis may only emerge after weeks of epigenetic adaptation [225]. The proposed beneficial effects of the KD may be mediated by a shift in the NAD+/NADH ratio [226], giving insight into how it could modulate lactate concentration. Further understanding the dynamic brain lactate response acutely and over time to any proposed nutritional or dietary protocol may provide valuable insight that explains how these approaches alter brain neurometabolism with relevance for psychiatric disorders.
  • Stress, hypoxia, and inflammation
Oxidative stress and inflammation impair mitochondrial function and result in excessive ROS production [227]. Acute and intermittent lactate exposure is adaptive [228] and can promote oxidative stress resistance, upregulating cellular defence mechanisms via NRF2 [229]. However, chronic exposure to lactate alters fatty acid metabolism and membrane composition, increasing ROS production and decreasing ATP production in mouse cardiomyocytes, resulting in further impairment of mitochondrial function [120]. Hypoxic and pseudohypoxic (NAD+ depleted) states lead to the stabilisation of Hypoxia Inducible Factor 1-alpha (HIF1-α) downstream of SIRT1 that disrupts mitochondrial function through genome desynchrony [230]. It almost goes without saying that hyperbaric oxygen therapy (HbO2) leads to a reduction in lactate concentration in skeletal muscle [231]. Cyanide leads to an accumulation of lactate in the rat brain that can be measured with microdialysis and ameliorated with HbO2. On the other hand, some evidence suggests that acute hypoxia may transiently improve metabolic parameters [232], and the lactate response to this may be an important variable to consider. HbO2 or acute hypoxia may represent novel treatment avenues for mood and affective disorders that deserve investigation tailored to the individual based on changes in lactate metabolism.
  • Circadian rhythm
Circadian biology is a key consideration for improving the health of mitochondria. This relationship is biochemically underpinned by melatonin, a mitochondrially synthesised hormone controlled by light that enhances electron transport, ATP synthesis, and ROS [233]. Knowing that lactate is associated with the sleep–wake cycle, and that a chronic elevation of lactate can result in excessive ROS production and mitochondrial dysfunction [120], addressing this axis with patients is then critical. This link enables the clinician to go beyond classical sleep hygiene measures and towards metabolic therapy for improving conditions such as insomnia.
  • Other examples of lactate manipulation
Lactate can be elevated via intravenous infusion of sodium lactate solution, documented to exhibit panicogenic and anti-depressive effects [42,66]. In a model of SCZ, exogenous lactate rescued the abnormal behaviour of DISC transgenic mice [234]. Early work discovered that repeated infusion led to a reduction in the anxiogenic response in participants over time, likely reflecting habituation to the interoceptive signals associated with elevated lactate [235]. Electroconvulsive therapy (ECT) is currently employed in psychiatry, and seizure activity is associated with elevations in circulating lactate and is elevated above 4 mM in those who suffer with post-ECT agitation but not in those who do not [236]. Whole-body hyperthermia has similar effects [237]. Several psychiatric medications influence the catecholamine system through their mechanism of action, and therefore, one would expect this to have an impact on lactate production, a worthwhile consideration for future work.

5. Brain Lactate Measurement with Functional Magnetic Resonance Spectroscopy

Proton magnetic resonance spectroscopy (1H-MRS) can detect and quantify lactate in brain tissue without administration of an exogenous tracer. Lactate exhibits a distinct peak with a chemical shift around 1.3 parts per million (ppm) within the spectra. This well-separated position from other metabolite resonances appears as a doublet due to the coupling between its two proton environments that can reflect the biochemical or biophysical properties in the environment.
The in vivo detection of lactate with 1H-MRS yields challenges. The lactate resonance has a relatively low signal strength compared to other metabolites and spectral overlap from compounds such as lipids and other macromolecules require careful analysis and spectral fitting to distinguish and quantify it. Shimming reduces inhomogeneity in magnetic fields and ensures that lactate appears at the expected 1.3 ppm spectral position. Various acquisition techniques can be employed, with the choice of echo time (TE) playing a crucial role. Short TE approaches are used to detect lactate minimizing the decrease in signal due to T2 relaxation, although these may increase the contribution from macromolecules. Long TE approaches reduce contributions from broad resonances but may decrease sensitivity to lactate. J-editing is an approach that selectively detects lactate by exploiting scalar coupling between the three protons on the C2 position and the proton on the C1 position in the lactate molecule. This approach suppresses unwanted signals and thereby enhances the lactate resonance. The drawback of lactate editing is a 40% reduction in SNR, compared to a non-edited acquisition with the same number of spectral averages. The choice of acquisition parameters, including TE, spectral editing techniques, and pulse sequences, is likely to depend on the MR field strength. It goes without saying that higher field strengths provide increased SNR and improved spectral separation, thus enhancing lactate detection.
Early work identified that photic stimulation of the visual cortex precipitates a rise in lactate concentration to 0.3–0.9 mM [74]. The visual stimulation paradigm effectively enables the quantification of lactate and other metabolites in the visual cortex due to high energy metabolism and proximity of the surface coil. Notably, it is in the visual cortex that differential lactate responses in panic disorder participants have been observed [88]. Other groups have been able to monitor a range of energy metabolites including lactate, ATP, and phosphocreatine (PCr) levels, shedding light on the metabolic demands of the motor cortex during movement [238]. However, although robust and replicated evidence reported altered lactate levels in areas of the brain relevant to mood and affective disorders, such as the anterior cingulate cortex, very few studies have aimed to quantify dynamic lactate responses in those areas with functional task-based paradigms [239].
In animal studies, microdialysis can be employed as a complementary technique to validate the MRS-derived signals [240]. This technique is challenging for many reasons, particularly because lactate responds to anaesthesia or immobilisation and other forms of stress, further highlighting the nature of lactate as a moment-to-moment biomarker of neuroenergetics.

6. Conclusions

A growing body of evidence already highlights the importance of cerebral lactate measurement in mood and affective disorders. Recent meta-analyses of both BPD and SCZ have identified alterations in brain lactate [31,241]. BPD in particular is strongly associated with evidence of mitochondrial dysfunction [32], and it appears that brain lactate can be modified by current pharmacological therapies in these patients [123,124]. We hope that by placing these data within a framework that enables a thorough understanding of their nuance, we can drive the development of better technology and research paradigms required to make precise and useful measurements [242,243].
In the present review, we have outlined roles for lactate that map onto each domain of the National Institutes of Mental Health RDoC framework (Figure 1) [34]. With data from both rodent and human studies, lactate in the brain is panicogenic [244] and anti-depressive [42], involved in social threat perception [245], memory formation and executive function [80,246], tied to circadian rhythms and arousal [62,78], and reflects motor activity [102]. Simple questions we can ask for future work relating to the field of metabolic psychiatry would be as follows: Where and when is lactate elevated? How long does it persist? And why?
Any therapeutic intervention should have both objective and subjective outcomes that allow a clinician and patient to monitor and validate progress. If we reconsider the seminal observations of abnormal lactate responses to exercise in anxious populations from our updated metabolic perspective, then we should see this as an opportunity for intervention that can be personalised and monitored over time. This has implications for drug development, mitochondrially targeted nutraceuticals or dietary interventions, and for developing exercise protocols that can modulate the capacity to produce or use lactate, influencing the temporal dynamics.
Overall, we posit that lactate connects metabolic stress at the cellular level to stress at the macroscopic level and, as such, naturally presents itself as a theranostic biomarker for furthering the field of metabolic psychiatry.

Author Contributions

Conceptualisation, E.C. and J.P.; writing—original draft preparation, E.C. and J.R.; writing—review and editing, A.C., I.R., J.P. and J.R.; supervision, A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, Y.; Zhao, W.; Lu, Y.; Zhao, Y.; Zhang, Y.; Dai, M.; Hai, S.; Ge, N.; Zhang, S.; Huang, M.; et al. Systematic metabolic characterization of mental disorders reveals age-related metabolic disturbances as potential risk factors for depression in older adults. Medcomm 2022, 3, e165. [Google Scholar] [CrossRef] [PubMed]
  2. Aguilar-López, B.A.; Moreno-Altamirano, M.M.B.; Dockrell, H.M.; Duchen, M.R.; Sánchez-García, F.J. Mitochondria: An Integrative Hub Coordinating Circadian Rhythms, Metabolism, the Microbiome, and Immunity. Front. Cell Dev. Biol. 2020, 8, 51. [Google Scholar] [CrossRef] [PubMed]
  3. San-Millán, I. The Key Role of Mitochondrial Function in Health and Disease. Antioxidants 2023, 12, 782. [Google Scholar] [CrossRef] [PubMed]
  4. Kann, O. The interneuron energy hypothesis: Implications for brain disease. Neurobiol. Dis. 2016, 90, 75–85. [Google Scholar] [CrossRef]
  5. Colasanti, A.; Bugiardini, E.; Amawi, S.; Poole, O.V.; Skorupinska, I.; Skorupinska, M.; Germain, L.; Kozyra, D.; Holmes, S.; James, N.; et al. Primary mitochondrial diseases increase susceptibility to bipolar affective disorder. J. Neurol. Neurosurg. Psychiatry 2020, 91, 892–894. [Google Scholar] [CrossRef]
  6. Onyango, I.G.; Lu, J.; Rodova, M.; Lezi, E.; Crafter, A.B.; Swerdlow, R.H. Regulation of neuron mitochondrial biogenesis and relevance to brain health. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2010, 1802, 228–234. [Google Scholar] [CrossRef]
  7. Picard, M.; McEwen, B.S. Psychological Stress and Mitochondria: A Systematic Review. Psychosom. Med. 2018, 80, 141–153. [Google Scholar] [CrossRef]
  8. Andreazza, A.C.; Nierenberg, A.A. Mitochondrial Dysfunction: At the Core of Psychiatric Disorders? Biol. Psychiatry 2018, 83, 718–719. [Google Scholar] [CrossRef]
  9. Giménez-Palomo, A.; Dodd, S.; Anmella, G.; Carvalho, A.F.; Scaini, G.; Quevedo, J.; Pacchiarotti, I.; Vieta, E.; Berk, M. The Role of Mitochondria in Mood Disorders: From Physiology to Pathophysiology and to Treatment. Front. Psychiatry 2021, 12, 546801. [Google Scholar] [CrossRef]
  10. Pinna, A.; Colasanti, A. The Neurometabolic Basis of Mood Instability: The Parvalbumin Interneuron Link—A Systematic Review and Meta-Analysis. Front. Pharmacol. 2021, 12, 689473. [Google Scholar] [CrossRef]
  11. Gong, Y.; Chai, Y.; Ding, J.-H.; Sun, X.-L.; Hu, G. Chronic mild stress damages mitochondrial ultrastructure and function in mouse brain. Neurosci. Lett. 2011, 488, 76–80. [Google Scholar] [CrossRef] [PubMed]
  12. Detka, J.; Kurek, A.; Kucharczyk, M.; Głombik, K.; Basta-Kaim, A.; Kubera, M.; Lasoń, W.; Budziszewska, B. Brain glucose metabolism in an animal model of depression. Neuroscience 2015, 295, 198–208. [Google Scholar] [CrossRef] [PubMed]
  13. Ernst, J.; Hock, A.; Henning, A.; Seifritz, E.; Boeker, H.; Grimm, S. Increased pregenual anterior cingulate glucose and lactate concentrations in major depressive disorder. Mol. Psychiatry 2016, 22, 113–119. [Google Scholar] [CrossRef]
  14. Bradley, K.; Mao, X.; Case, J.; Kang, G.; Shungu, D.; Gabbay, V. Increased ventricular cerebrospinal fluid lactate in depressed adolescents. Eur. Psychiatry 2016, 32, 1–8. [Google Scholar] [CrossRef] [PubMed]
  15. Regenold, W.T.; Phatak, P.; Marano, C.M.; Sassan, A.; Conley, R.R.; Kling, M.A. Elevated cerebrospinal fluid lactate concentrations in patients with bipolar disorder and schizophrenia: Implications for the mitochondrial dysfunction hypothesis. Biol. Psychiatry 2009, 65, 489–494. [Google Scholar] [CrossRef]
  16. Millar, J.K.; Christie, S.; Anderson, S.; Lawson, D.; Loh, D.; Devon, R.S.; Arveiler, B.; Muir, W.J.; Blackwood, D.H.; Porteous, D.J. Genomic structure and localisation within a linkage hotspot of Disrupted In Schizophrenia 1, a gene disrupted by a translocation segregating with schizophrenia. Mol. Psychiatry 2001, 6, 173–178. [Google Scholar] [CrossRef]
  17. Jouroukhin, Y.; Kageyama, Y.; Misheneva, V.; Shevelkin, A.; Andrabi, S.; Prandovszky, E.; Yolken, R.H.; Dawson, V.L.; Dawson, T.M.; Aja, S.; et al. DISC1 regulates lactate metabolism in astrocytes: Implications for psychiatric disorders. Transl. Psychiatry 2018, 8, 1–12. [Google Scholar] [CrossRef]
  18. Sullivan, C.R.; O’donovan, S.M.; McCullumsmith, R.E.; Ramsey, A. Defects in Bioenergetic Coupling in Schizophrenia. Biol. Psychiatry 2018, 83, 739–750. [Google Scholar] [CrossRef]
  19. Roberts, R.; Barksdale, K.; Roche, J.; Lahti, A. Decreased synaptic and mitochondrial density in the postmortem anterior cingulate cortex in schizophrenia. Schizophr. Res. 2015, 168, 543–553. [Google Scholar] [CrossRef]
  20. Du, F.; Cooper, A.J.; Thida, T.; Sehovic, S.; Lukas, S.E.; Cohen, B.M.; Zhang, X.; Öngür, D. In Vivo Evidence for Cerebral Bioenergetic Abnormalities in Schizophrenia Measured Using31P Magnetization Transfer Spectroscopy. JAMA Psychiatry 2014, 71, 19–27. [Google Scholar] [CrossRef]
  21. Rowland, L.M.; Pradhan, S.; Korenic, S.; Wijtenburg, S.A.; Hong, L.E.; Edden, R.A.; Barker, P.B. Elevated brain lactate in schizophrenia: A 7 T magnetic resonance spectroscopy study. Transl. Psychiatry 2016, 6, e967. [Google Scholar] [CrossRef]
  22. Fizíková, I.; Dragašek, J.; Račay, P. Mitochondrial Dysfunction, Altered Mitochondrial Oxygen, and Energy Metabolism Associated with the Pathogenesis of Schizophrenia. Int. J. Mol. Sci. 2023, 24, 7991. [Google Scholar] [CrossRef] [PubMed]
  23. Cataldo, A.M.; McPhie, D.L.; Lange, N.T.; Punzell, S.; Elmiligy, S.; Ye, N.Z.; Froimowitz, M.P.; Hassinger, L.C.; Menesale, E.B.; Sargent, L.W.; et al. Abnormalities in Mitochondrial Structure in Cells from Patients with Bipolar Disorder. Am. J. Pathol. 2010, 177, 575–585. [Google Scholar] [CrossRef] [PubMed]
  24. Andreazza, A.C.; Shao, L.; Wang, J.-F.; Young, L.T. Mitochondrial Complex I Activity and Oxidative Damage to Mitochondrial Proteins in the Prefrontal Cortex of Patients With Bipolar Disorder. Arch. Gen. Psychiatry 2010, 67, 360–368. [Google Scholar] [CrossRef] [PubMed]
  25. Andreazza, A.C.; Wang, J.-F.; Salmasi, F.; Shao, L.; Young, L.T. Specific subcellular changes in oxidative stress in prefrontal cortex from patients with bipolar disorder. J. Neurochem. 2013, 127, 552–561. [Google Scholar] [CrossRef]
  26. CKonradi, C.; Eaton, M.; MacDonald, M.L.; Walsh, J.; Benes, F.M.; Heckers, S. Molecular Evidence for Mitochondrial Dysfunction in Bipolar Disorder. Arch. Gen. Psychiatry 2004, 61, 300–308. [Google Scholar] [CrossRef]
  27. Bodenstein, D.F.; Kim, H.K.; Brown, N.C.; Navaid, B.; Young, L.T.; Andreazza, A.C. Mitochondrial DNA content and oxidation in bipolar disorder and its role across brain regions. NPJ Schizophr. 2019, 5, 1–8. [Google Scholar] [CrossRef]
  28. Kato, T.; Takahashi, S.; Shioiri, T.; Murashita, J.; Hamakawa, H.; Inubushi, T. Reduction of brain phosphocreatine in bipolar II disorder detected by phosphorus-31 magnetic resonance spectroscopy. J. Affect. Disord. 1994, 31, 125–133. [Google Scholar] [CrossRef]
  29. Du, F.; Yuksel, C.; Chouinard, V.-A.; Huynh, P.; Ryan, K.; Cohen, B.M.; Öngür, D. Abnormalities in High-Energy Phosphate Metabolism in First-Episode Bipolar Disorder Measured Using 31P-Magnetic Resonance Spectroscopy. Biol. Psychiatry 2017, 84, 797–802. [Google Scholar] [CrossRef]
  30. Yuksel, C.; Du, F.; Ravichandran, C.; Goldbach, J.R.; Thida, T.; Lin, P.; Dora, B.; Gelda, J.; O’Connor, L.; Sehovic, S.; et al. Abnormal high-energy phosphate molecule metabolism during regional brain activation in patients with bipolar disorder. Mol. Psychiatry 2015, 20, 1079–1084. [Google Scholar] [CrossRef]
  31. Kuang, H.; Duong, A.; Jeong, H.; Zachos, K.; Andreazza, A.C. Lactate in bipolar disorder: A systematic review and meta-analysis. Psychiatry Clin. Neurosci. 2018, 72, 546–555. [Google Scholar] [CrossRef] [PubMed]
  32. Stork, C.; Renshaw, P.F. Mitochondrial dysfunction in bipolar disorder: Evidence from magnetic resonance spectroscopy research. Mol. Psychiatry 2005, 10, 900–919. [Google Scholar] [CrossRef] [PubMed]
  33. BCuthbert, B.N. The RDoC framework: Facilitating transition from ICD/DSM to dimensional approaches that integrate neuroscience and psychopathology. World Psychiatry 2014, 13, 28–35. [Google Scholar] [CrossRef] [PubMed]
  34. Definitions of the RDoC Domains and Constructs. National Institute of Mental Health (NIMH). Available online: https://www.nimh.nih.gov/research/research-funded-by-nimh/rdoc/definitions-of-the-rdoc-domains-and-constructs (accessed on 18 May 2023).
  35. San-Millán, I.; Brooks, G.A. Assessment of Metabolic Flexibility by Means of Measuring Blood Lactate, Fat, and Carbohydrate Oxidation Responses to Exercise in Professional Endurance Athletes and Less-Fit Individuals. Sports Med. 2018, 48, 467–479. [Google Scholar] [CrossRef] [PubMed]
  36. Poole, D.C.; Rossiter, H.B.; Brooks, G.A.; Gladden, L.B. The anaerobic threshold: 50+ years of controversy. J. Physiol. 2020, 599, 737–767. [Google Scholar]
  37. Cohen, M.E.; White, P.D. Life Situations, Emotions, and Neurocirculatory Asthenia (Anxiety Neurosis, Neurasthenia, Effort Syndrome). Psychosom. Med. 1951, 13, 335–357. [Google Scholar] [CrossRef]
  38. Holmgren, A.; Strom, G. Blood lactate concentration in relation to absolute and relative work load in normal men, and in mitral stenosis, atrial septal defect and vasoregulatory asthenia. Acta Med. Scand. 1959, 163, 185–193. [Google Scholar] [CrossRef]
  39. Jones, M.; Mellersh, V.; Musgrave, M. A Comparison of the Exercise Response in Anxiety States and Normal Controls. Psychosom. Med. 1946, 8, 180–187. [Google Scholar] [CrossRef]
  40. Linko, E. Lactic acid response to muscular exercise in neurocirculatory asthenia. Ann. Med. Intern. Fenn. 1950, 39, 161–176. [Google Scholar]
  41. Pitts, F.N.; McClure, J.N. Lactate Metabolism in Anxiety Neurosis. N. Engl. J. Med. 1967, 277, 1329–1336. [Google Scholar] [CrossRef]
  42. Carrard, A.; Elsayed, M.; Margineanu, M.; Boury-Jamot, B.; Fragnière, L.; Meylan, E.M.; Petit, J.M.; Fiumelli, H.; Magistretti, P.J.; Martin, J.L. Peripheral administration of lactate produces antidepressant-like effects. Mol. Psychiatry 2018, 23, 392–399. [Google Scholar] [CrossRef] [PubMed]
  43. Schurr, A. From rags to riches: Lactate ascension as a pivotal metabolite in neuroenergetics. Front. Neurosci. 2023, 17, 1145358. [Google Scholar] [CrossRef] [PubMed]
  44. Rabinowitz, J.D.; Enerbäck, S. Lactate: The ugly duckling of energy metabolism. Nat. Metab. 2020, 2, 566–571. [Google Scholar] [CrossRef] [PubMed]
  45. Brooks, G.A.; Arevalo, J.A.; Osmond, A.D.; Leija, R.G.; Curl, C.C.; Tovar, A.P. Lactate in contemporary biology: A phoenix risen. J. Physiol. 2022, 600, 1229–1251. [Google Scholar]
  46. Aryaman, J.; Johnston, I.G.; Jones, N.S. Mitochondrial Heterogeneity. Front. Genet. 2019, 9, 718. [Google Scholar]
  47. Wallace, D.C.; Chalkia, D. Mitochondrial DNA Genetics and the Heteroplasmy Conundrum in Evolution and Disease. Cold Spring Harb. Perspect. Biol. 2013, 5, a021220. [Google Scholar] [CrossRef]
  48. Maddock, R.J.; Casazza, G.A.; Buonocore, M.H.; Tanase, C. Vigorous exercise increases brain lactate and Glx (glutamate+glutamine): A dynamic 1H-MRS study. NeuroImage 2011, 57, 1324–1330. [Google Scholar] [CrossRef]
  49. Shirbandi, K.B.; Rikhtegar, R.; Khalafi, M.; Attari, M.M.A.; Rahmani, F.; Javanmardi, P.B.; Iraji, S.M.; Aghdam, Z.B.; Rashnoudi, A.M.R. Functional Magnetic Resonance Spectroscopy of Lactate in Alzheimer Disease: A Comprehensive Review of Alzheimer Disease Pathology and the Role of Lactate. Top. Magn. Reson. Imaging 2023, 32, 15–26. [Google Scholar]
  50. Gladden, L.B. Lactate metabolism: A new paradigm for the third millennium. J. Physiol. 2004, 558 Pt 1, 5–30. [Google Scholar]
  51. Gibbs, M.E.; Hertz, L. Inhibition of astrocytic energy metabolism by d-lactate exposure impairs memory. Neurochem. Int. 2008, 52, 1012–1018. [Google Scholar] [CrossRef]
  52. Li, X.; Zhang, Y.; Xu, L.; Wang, A.; Zou, Y.; Li, T.; Huang, L.; Chen, W.; Liu, S.; Jiang, K.; et al. Ultrasensitive sensors reveal the spatiotemporal landscape of lactate metabolism in physiology and disease. Cell Metab. 2023, 35, 200–211.e9. [Google Scholar] [CrossRef] [PubMed]
  53. Brooks, G.A. Lactate as a fulcrum of metabolism. Redox Biol. 2020, 35, 101454. [Google Scholar] [CrossRef] [PubMed]
  54. Hui, S.; Ghergurovich, J.M.; Morscher, R.J.; Jang, C.; Teng, X.; Lu, W.; Esparza, L.A.; Reya, T.; Zhan, L.; Guo, J.Y.; et al. Glucose feeds the TCA cycle via circulating lactate. Nature 2017, 551, 115–118. [Google Scholar] [CrossRef]
  55. Hui, S.; Cowan, A.J.; Zeng, X.; Yang, L.; TeSlaa, T.; Li, X.; Bartman, C.; Zhang, Z.; Jang, C.; Wang, L.; et al. Quantitative Fluxomics of Circulating Metabolites. Cell Metab. 2020, 32, 676–688.e4. [Google Scholar] [CrossRef]
  56. Parikh, S.; Goldstein, A.; Koenig, M.K.; Scaglia, F.; Enns, G.M.; Saneto, R.; Anselm, I.; Cohen, B.H.; Falk, M.J.; Greene, C.; et al. Diagnosis and management of mitochondrial disease: A consensus statement from the Mitochondrial Medicine Society. Anesth. Analg. 2015, 17, 689–701. [Google Scholar] [CrossRef] [PubMed]
  57. Jones, T.E.; Pories, W.J.; Houmard, J.A.; Tanner, C.J.; Zheng, D.; Zou, K.; Coen, P.M.; Goodpaster, B.H.; Kraus, W.E.; Dohm, G.L. Plasma lactate as a marker of metabolic health: Implications of elevated lactate for impairment of aerobic metabolism in the metabolic syndrome. Surgery 2019, 166, 861–866. [Google Scholar] [CrossRef] [PubMed]
  58. Bergman, B.C.; Wolfel, E.E.; Butterfield, G.E.; Lopaschuk, G.D.; Casazza, G.A.; Horning, M.A.; Brooks, G.A.; Sun, S.; Li, H.; Chen, J.; et al. Active muscle and whole body lactate kinetics after endurance training in men. J. Appl. Physiol. 1999, 87, 1684–1696. [Google Scholar] [CrossRef]
  59. Donovan, C.M.; Brooks, G.A. Endurance training affects lactate clearance, not lactate production. Am. J. Physiol.-Endocrinol. Metab. 1983, 244, E83–E92. [Google Scholar] [CrossRef]
  60. Wallace, N.K.; Pollard, F.; Savenkova, M.; Karatsoreos, I.N. Effect of aging on daily rhythms in lactate metabolism in the medial prefrontal cortex of male mice. Neuroscience 2020, 448, 300–310. [Google Scholar] [CrossRef]
  61. Shram, N.; Netchiporouk, L.; Cespuglio, R. Lactate in the brain of the freely moving rat: Voltammetric monitoring of the changes related to the sleep-wake states. Eur. J. Neurosci. 2002, 16, 461–466. [Google Scholar] [CrossRef]
  62. Lundgaard, I.; Lu, M.L.; Yang, E.; Peng, W.; Mestre, H.; Hitomi, E.; Deane, R.; Nedergaard, M. Glymphatic clearance controls state-dependent changes in brain lactate concentration. J. Cereb. Blood Flow Metab. 2016, 37, 2112–2124. [Google Scholar] [CrossRef]
  63. Noort, S.V.D.; Brine, K. Effect of sleep on brain labile phosphates and metabolic rate. Am. J. Physiol. Content 1970, 218, 1434–1439. [Google Scholar] [CrossRef] [PubMed]
  64. Wisor, J.P.; Rempe, M.J.; Schmidt, M.A.; Moore, M.E.; Clegern, W.C. Sleep Slow-Wave Activity Regulates Cerebral Glycolytic Metabolism. Cereb. Cortex 2012, 23, 1978–1987. [Google Scholar] [CrossRef] [PubMed]
  65. Hadjihambi, A.; Karagiannis, A.; Theparambil, S.M.; Ackland, G.L.; Gourine, A.V. The effect of general anaesthetics on brain lactate release. Eur. J. Pharmacol. 2020, 881, 173188. [Google Scholar] [CrossRef] [PubMed]
  66. Koenigsberg, H.W.; Pollak, C.P.; Fine, J.; Kakuma, T. Lactate sensitivity in sleeping panic disorder patients and healthy controls. Biol. Psychiatry 1992, 32, 539–542. [Google Scholar] [CrossRef] [PubMed]
  67. Naylor, E.; Aillon, D.V.; Barrett, B.S.; Wilson, G.S.; Johnson, D.A.; Johnson, D.A.; Harmon, H.P.; Gabbert, S.; Petillo, P.A. Lactate as a Biomarker for Sleep. Sleep 2012, 35, 1209–1222. [Google Scholar] [CrossRef]
  68. Ramanathan, L.; Hu, S.; Frautschy, S.A.; Siegel, J.M. Short-term total sleep deprivation in the rat increases antioxidant responses in multiple brain regions without impairing spontaneous alternation behavior. Behav. Brain Res. 2010, 207, 305–309. [Google Scholar] [CrossRef]
  69. Boumezbeur, F.; Petersen, K.F.; Cline, G.W.; Mason, G.F.; Behar, K.L.; Shulman, G.I.; Rothman, D.L. The Contribution of Blood Lactate to Brain Energy Metabolism in Humans Measured by Dynamic13C Nuclear Magnetic Resonance Spectroscopy. J. Neurosci. 2010, 30, 13983–13991. [Google Scholar] [CrossRef]
  70. Díaz-García, C.M.; Yellen, G. Neurons rely on glucose rather than astrocytic lactate during stimulation. J. Neurosci. Res. 2018, 97, 883–889. [Google Scholar] [CrossRef]
  71. Brown, A.M.; Ransom, B.R. Astrocyte glycogen and brain energy metabolism. Glia 2007, 55, 1263–1271. [Google Scholar] [CrossRef]
  72. Hu, Y.; Wilson, G.S. A Temporary Local Energy Pool Coupled to Neuronal Activity: Fluctuations of Extracellular Lactate Levels in Rat Brain Monitored with Rapid-Response Enzyme-Based Sensor. J. Neurochem. 1997, 69, 1484–1490. [Google Scholar] [CrossRef] [PubMed]
  73. Magistretti, P.J.; Allaman, I. Lactate in the brain: From metabolic end-product to signalling molecule. Nat. Rev. Neurosci. 2018, 19, 235–249. [Google Scholar] [PubMed]
  74. Prichard, J.; Rothman, D.; Novotny, E.; Petroff, O.; Kuwabara, T.; Avison, M.; Howseman, A.; Hanstock, C.; Shulman, R. Lactate rise detected by 1H NMR in human visual cortex during physiologic stimulation. Proc. Natl. Acad. Sci. USA 1991, 88, 5829–5831. [Google Scholar] [CrossRef] [PubMed]
  75. Vanzetta, I.; Grinvald, A. Increased Cortical Oxidative Metabolism Due to Sensory Stimulation: Implications for Functional Brain Imaging. Science 1999, 286, 1555–1558. [Google Scholar] [CrossRef] [PubMed]
  76. Pellerin, L.; Pellegri, G.; Bittar, P.G.; Charnay, Y.; Bouras, C.; Martin, J.-L.; Stella, N.; Magistretti, P.J. Evidence Supporting the Existence of an Activity-Dependent Astrocyte-Neuron Lactate Shuttle. Dev. Neurosci. 1998, 20, 291–299. [Google Scholar] [CrossRef]
  77. Schurr, A.; Miller, J.J.; Payne, R.S.; Rigor, B.M. An Increase in Lactate Output by Brain Tissue Serves to Meet the Energy Needs of Glutamate-Activated Neurons. J. Neurosci. 1999, 19, 34–39. [Google Scholar] [CrossRef]
  78. Zuend, M.; Saab, A.S.; Wyss, M.T.; Ferrari, K.D.; Hösli, L.; Looser, Z.J.; Stobart, J.L.; Duran, J.; Guinovart, J.J.; Barros, L.F.; et al. Arousal-induced cortical activity triggers lactate release from astrocytes. Nat. Metab. 2020, 2, 179–191. [Google Scholar] [CrossRef]
  79. Elekesa, O.; Venema, K.; Postema, F.; Dringen, R.; Hamprecht, B.; Korf, J. Evidence that stress activates glial lactate formation in vivo assessed with rat hippocampus lactography. Neurosci. Lett. 1996, 208, 69–72. [Google Scholar] [CrossRef]
  80. Suzuki, A.; Stern, S.A.; Bozdagi, O.; Huntley, G.W.; Walker, R.H.; Magistretti, P.J.; Alberini, C.M. Astrocyte-Neuron Lactate Transport Is Required for Long-Term Memory Formation. Cell 2011, 144, 810–823. [Google Scholar] [CrossRef]
  81. Newman, L.A.; Korol, D.L.; Gold, P.E. Lactate Produced by Glycogenolysis in Astrocytes Regulates Memory Processing. PLoS ONE 2011, 6, e28427. [Google Scholar] [CrossRef]
  82. Murphy-Royal, C.; Johnston, A.D.; Boyce, A.K.J.; Diaz-Castro, B.; Institoris, A.; Peringod, G.; Zhang, O.; Stout, R.F.; Spray, D.C.; Thompson, R.J.; et al. Stress gates an astrocytic energy reservoir to impair synaptic plasticity. Nat. Commun. 2020, 11, 2014. [Google Scholar] [CrossRef] [PubMed]
  83. Zhao, Y.; Zhang, Q.; Shao, X.; Ouyang, L.; Wang, X.; Zhu, K.; Chen, L. Decreased Glycogen Content Might Contribute to Chronic Stress-Induced Atrophy of Hippocampal Astrocyte volume and Depression-like Behavior in Rats. Sci. Rep. 2017, 7, 43192. [Google Scholar] [CrossRef] [PubMed]
  84. Quistorff, B.; Secher, N.H.; Van Lieshout, J.J. Lactate fuels the human brain during exercise. FASEB J. 2008, 22, 3443–3449. [Google Scholar] [CrossRef]
  85. Kemppainen, J.; Aalto, S.; Fujimoto, T.; Kalliokoski, K.K.; Långsjö, J.; Oikonen, V.; Rinne, J.; Nuutila, P.; Knuuti, J. High intensity exercise decreases global brain glucose uptake in humans. J. Physiol. 2005, 568, 323–332. [Google Scholar] [CrossRef]
  86. Van Hall, G. Lactate kinetics in human tissues at rest and during exercise. Acta Physiol. 2010, 199, 499–508. [Google Scholar] [CrossRef] [PubMed]
  87. Barros, L.F. Metabolic signaling by lactate in the brain. Trends Neurosci. 2013, 36, 396–404. [Google Scholar] [CrossRef]
  88. Maddock, R.J.; Buonocore, M.H.; Copeland, L.E.; Richards, A.L. Elevated brain lactate responses to neural activation in panic disorder: A dynamic 1H-MRS study. Mol. Psychiatry 2008, 14, 537–545. [Google Scholar] [CrossRef]
  89. Dager, S.R.; Yim, J.B.; Khalil, G.E.; Artru, A.A.; Bowden, D.M.; Kenny, M.A. Application of a Novel Fiber-Optic Biosensor In Situ to Investigate the Metabolic Effect of Lactate Infusion. Neuropsychopharmacology 1995, 12, 307–313. [Google Scholar] [CrossRef]
  90. Garcia, C.K.; Goldstein, J.L.; Pathak, R.K.; Anderson, R.G.W.; Brown, M.S. Molecular characterization of a membrane transporter for lactate, pyruvate, and other monocarboxylates: Implications for the Cori cycle. Cell 1994, 76, 865–873. [Google Scholar] [CrossRef]
  91. Sinning, A.; Hübner, C.A. Minireview: pH and synaptic transmission. FEBS Lett. 2013, 587, 1923–1928. [Google Scholar] [CrossRef]
  92. Waldmann, R.; Champigny, G.; Bassilana, F.; Heurteaux, C.; Lazdunski, M. A proton-gated cation channel involved in acid-sensing. Nature 1997, 386, 173–177. [Google Scholar] [CrossRef] [PubMed]
  93. Esquivel, G.; Schruers, K.R.; Maddock, R.J.; Colasanti, A.; Griez, E.J. Acids in the brain: A factor in panic? J. Psychopharmacol. 2010, 24, 639–647. [Google Scholar] [CrossRef] [PubMed]
  94. Milbank, E.; López, M. Orexins/Hypocretins: Key Regulators of Energy Homeostasis. Front. Endocrinol. 2019, 10, 830. [Google Scholar] [CrossRef] [PubMed]
  95. Date, Y.; Ueta, Y.; Yamashita, H.; Yamaguchi, H.; Matsukura, S.; Kangawa, K.; Sakurai, T.; Yanagisawa, M.; Nakazato, M. Orexins, orexigenic hypothalamic peptides, interact with autonomic, neuroendocrine and neuroregulatory systems’. Proc. Natl. Acad. Sci. USA 1999, 96, 748–753. [Google Scholar] [CrossRef]
  96. Severson, C.A.; Wang, W.; Pieribone, V.A.; Dohle, C.I.; Richerson, G.B. Midbrain serotonergic neurons are central pH chemoreceptors. Nat. Neurosci. 2003, 6, 1139–1140. [Google Scholar] [CrossRef]
  97. Gargaglioni, L.H.; Hartzler, L.K.; Putnam, R.W. The locus coeruleus and central chemosensitivity. Respir. Physiol. Neurobiol. 2010, 173, 264–273. [Google Scholar] [CrossRef]
  98. Lauritzen, K.H.; Morland, C.; Puchades, M.; Holm-Hansen, S.; Hagelin, E.M.; Lauritzen, F.; Attramadal, H.; Storm-Mathisen, J.; Gjedde, A.; Bergersen, L.H. Lactate receptor sites link neurotransmission, neurovascular coupling, and brain energy metabolism. Cereb. Cortex 2014, 24, 2784–2795. [Google Scholar] [CrossRef]
  99. Morland, C.; Lauritzen, K.H.; Puchades, M.; Holm-Hansen, S.; Andersson, K.; Gjedde, A.; Attramadal, H.; Storm-Mathisen, J.; Bergersen, L.H. The lactate receptor, G-protein-coupled receptor 81/hydroxycarboxylic acid receptor 1: Expression and action in brain. J. Neurosci. Res. 2015, 93, 1045–1055. [Google Scholar] [CrossRef]
  100. De Castro Abrantes, H.; Briquet, M.; Schmuziger, C.; Restivo, L.; Puyal, J.; Rosenberg, N.; Rocher, A.B.; Offermanns, S.; Chatton, J.Y. The Lactate Receptor HCAR1 Modulates Neuronal Network Activity through the Activation of Gα and Gβγ Subunits. J. Neurosci. 2019, 39, 4422–4433. [Google Scholar] [CrossRef]
  101. Gilbert, E.; Tang, J.M.; Ludvig, N.; Bergold, P.J. Elevated lactate suppresses neuronal firing in vivo and inhibits glucose metabolism in hippocampal slice cultures. Brain Res. 2006, 1117, 213–223. [Google Scholar] [CrossRef]
  102. Coco, M.; Alagona, G.; Rapisarda, G.; Costanzo, E.; Calogero, R.A.; Perciavalle, V.; Perciavalle, V. Md Elevated blood lactate is associated with increased motor cortex excitability. Somatosens. Mot. Res. 2010, 27, 1–8. [Google Scholar] [CrossRef]
  103. Hollnagel, J.-O.; Cesetti, T.; Schneider, J.; Vazetdinova, A.; Valiullina-Rakhmatullina, F.; Lewen, A.; Rozov, A.; Kann, O. Lactate Attenuates Synaptic Transmission and Affects Brain Rhythms Featuring High Energy Expenditure. iScience 2020, 23, 101316. [Google Scholar] [CrossRef] [PubMed]
  104. Scavuzzo, C.J.; Rakotovao, I.; Dickson, C.T. Differential effects of L- and D-lactate on memory encoding and consolidation: Potential role of HCAR1 signaling. Neurobiol. Learn. Mem. 2020, 168, 107151. [Google Scholar] [CrossRef]
  105. Lev-Vachnish, Y.; Cadury, S.; Rotter-Maskowitz, A.; Feldman, N.; Roichman, A.; Illouz, T.; Varvak, A.; Nicola, R.; Madar, R.; Okun, E. L-Lactate Promotes Adult Hippocampal Neurogenesis. Front. Neurosci. 2019, 13, 403. [Google Scholar] [CrossRef]
  106. Lambertus, M.; Øverberg, L.T.; Andersson, K.A.; Hjelden, M.S.; Hadzic, A.; Haugen, Ø.P.; Storm-Mathisen, J.; Bergersen, L.H.; Geiseler, S.; Morland, C. L-lactate induces neurogenesis in the mouse ventricular-subventricular zone via the lactate receptor HCA1. Acta Physiol. 2021, 231, e13587. [Google Scholar]
  107. Lambertus, M.; Øverberg, L.T.; Andersson, K.A.; Hjelden, M.S.; Hadzic, A.; Haugen, Ø.P.; Storm-Mathisen, J.; Bergersen, L.H.; Geiseler, S.; Morland, C. Exercise induces cerebral VEGF and angiogenesis via the lactate receptor HCAR1. Nat. Commun. 2017, 8, 15557. [Google Scholar]
  108. Vardjan, N.; Chowdhury, H.H.; Horvat, A.; Velebit, J.; Malnar, M.; Muhič, M.; Kreft, M.; Krivec, Š.G.; Bobnar, S.T.; Miš, K.; et al. Enhancement of Astroglial Aerobic Glycolysis by Extracellular Lactate-Mediated Increase in cAMP. Front. Mol. Neurosci. 2018, 11, 148. [Google Scholar] [CrossRef] [PubMed]
  109. Tang, F.; Lane, S.; Korsak, A.; Paton, J.F.; Gourine, A.V.; Kasparov, S.; Teschemacher, A.G. Lactate-mediated glia-neuronal signalling in the mamalian brain. Nat. Commun. 2014, 5, 3284. [Google Scholar] [CrossRef]
  110. Shaif, N.A.; Jang, D.; Cho, D.; Kim, S.; Seo, D.B.; Shim, I. The Antidepressant-Like Effect of Lactate in an Animal Model of Menopausal Depression. Biomedicines 2018, 6, 108. [Google Scholar] [CrossRef]
  111. Gordon, G.R.J.; Choi, H.B.; Rungta, R.L.; Ellis-Davies, G.C.R.; MacVicar, B.A. Brain metabolism dictates the polarity of astrocyte control over arterioles. Nature 2008, 456, 745–749. [Google Scholar] [CrossRef]
  112. Mintun, M.A.; Vlassenko, A.G.; Rundle, M.M.; Raichle, M.E. Increased lactate/pyruvate ratio augments blood flow in physiologically activated human brain. Proc. Natl. Acad. Sci. USA 2004, 101, 659–664. [Google Scholar] [CrossRef] [PubMed]
  113. Wu, J.; Holstein, J.D.; Upadhyay, G.; Lin, D.-T.; Conway, S.; Muller, E.; Lechleiter, J.D. Purinergic Receptor-Stimulated IP3-Mediated Ca2+ Release Enhances Neuroprotection by Increasing Astrocyte Mitochondrial Metabolism during Aging. J. Neurosci. 2007, 27, 6510–6520. [Google Scholar] [CrossRef] [PubMed]
  114. Requardt, R.P.; Hirrlinger, P.G.; Wilhelm, F.; Winkler, U.; Besser, S.; Hirrlinger, J. Ca2+ signals of astrocytes are modulated by the NAD+/NADH redox state. J. Neurochem. 2012, 120, 1014–1025. [Google Scholar] [CrossRef]
  115. Parsons, M.P.; Hirasawa, M. ATP-Sensitive Potassium Channel-Mediated Lactate Effect on Orexin Neurons: Implications for Brain Energetics during Arousal. J. Neurosci. 2010, 30, 8061–8070. [Google Scholar] [CrossRef] [PubMed]
  116. Li, R.; Yang, Y.; Wang, H.; Zhang, T.; Duan, F.; Wu, K.; Yang, S.; Xu, K.; Jiang, X.; Sun, X. Lactate and Lactylation in the Brain: Current Progress and Perspectives. Cell Mol. Neurobiol. 2023, 43, 2541–2555. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, D.; Tang, Z.; Huang, H.; Zhou, G.; Cui, C.; Weng, Y.; Liu, W.; Kim, S.; Lee, S.; Perez-Neut, M.; et al. Metabolic regulation of gene expression by histone lactylation. Nature 2019, 574, 575–580. [Google Scholar] [CrossRef] [PubMed]
  118. Liu, X.; Zhang, Y.; Li, W.; Zhou, X. Lactylation, an emerging hallmark of metabolic reprogramming: Current progress and open challenges. Front. Cell Dev. Biol. 2022, 10, 972020. [Google Scholar] [CrossRef]
  119. Hagihara, H.; Shoji, H.; Otabi, H.; Toyoda, A.; Katoh, K.; Namihira, M.; Miyakawa, T. Protein lactylation induced by neural excitation. Cell Rep. 2021, 37, 109820. [Google Scholar] [CrossRef]
  120. San-Millan, I.; Sparagna, G.C.; Chapman, H.L.; Warkins, V.L.; Chatfield, K.C.; Shuff, S.R.; Martinez, J.L.; Brooks, G.A. Chronic Lactate Exposure Decreases Mitochondrial Function by Inhibition of Fatty Acid Uptake and Cardiolipin Alterations in Neonatal Rat Cardiomyocytes. Front. Nutr. 2022, 9, 809485. [Google Scholar] [CrossRef]
  121. Manzar, H.; Abdulhussein, D.; Yap, T.E.; Cordeiro, M.F. Cellular Consequences of Coenzyme Q10 Deficiency in Neurodegeneration of the Retina and Brain. Int. J. Mol. Sci. 2020, 21, 9299. [Google Scholar] [CrossRef]
  122. Parikh, S.; The Mitochondrial Medicine Society; Saneto, R.; Falk, M.J.; Anselm, I.; Cohen, B.H.; Haas, R. A modern approach to the treatment of mitochondrial disease. Curr. Treat. Options Neurol. 2009, 11, 414–430. [Google Scholar] [CrossRef] [PubMed]
  123. Machado-Vieira, R.; Zanetti, M.V.; Otaduy, M.C.; De Sousa, R.T.; Soeiro-de-Souza, M.G.; Costa, A.C.; Carvalho, A.F.; Leite, C.C.; Busatto, G.F.; Zarate, C.A., Jr.; et al. Increased Brain Lactate During Depressive Episodes and Reversal Effects by Lithium Monotherapy in Drug-Naive Bipolar Disorder: A 3-T 1H-MRS Study. J. Clin. Psychopharmacol. 2017, 37, 40–45. [Google Scholar] [CrossRef] [PubMed]
  124. Kim, D.J.; Lyoo, I.K.; Yoon, S.J.; Choi, T.; Lee, B.; Kim, J.E.; Lee, J.S.; Renshaw, P.F. Clinical response of quetiapine in rapid cycling manic bipolar patients and lactate level changes in proton magnetic resonance spectroscopy. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2007, 31, 1182–1188. [Google Scholar] [CrossRef] [PubMed]
  125. Brivio, P.; Audano, M.; Gallo, M.T.; Miceli, E.; Gruca, P.; Lason, M.; Litwa, E.; Fumagalli, F.; Papp, M.; Mitro, N.; et al. Venlafaxine’s effect on resilience to stress is associated with a shift in the balance between glucose and fatty acid utilization. Neuropsychopharmacology 2023, 48, 1475–1483. [Google Scholar] [CrossRef] [PubMed]
  126. Forester, B.P.M.; Harper, D.G.; Georgakas, J.; Ravichandran, C.; Madurai, N.; Cohen, B.M. Antidepressant Effects of Open Label Treatment With Coenzyme Q10 in Geriatric Bipolar Depression. J. Clin. Psychopharmacol. 2015, 35, 338–340. [Google Scholar] [CrossRef]
  127. Lee, S.; Choi, Y.; Jeong, E.; Park, J.; Kim, J.; Tanaka, M.; Choi, J. Physiological significance of elevated levels of lactate by exercise training in the brain and body. J. Biosci. Bioeng. 2023, 135, 167–175. [Google Scholar] [CrossRef]
  128. Park, J.; Kim, J.; Mikami, T. Exercise-Induced Lactate Release Mediates Mitochondrial Biogenesis in the Hippocampus of Mice via Monocarboxylate Transporters. Front. Physiol. 2021, 12, 736905. [Google Scholar] [CrossRef]
  129. El Hayek, L.; Khalifeh, M.; Zibara, V.; Abi Assaad, R.; Emmanuel, N.; Karnib, N.; El-Ghandour, R.; Nasrallah, P.; Bilen, M.; Ibrahim, P.; et al. Lactate Mediates the Effects of Exercise on Learning and Memory through SIRT1-Dependent Activation of Hippocampal Brain-Derived Neurotrophic Factor (BDNF). J. Neurosci. 2019, 39, 2369–2382. [Google Scholar] [CrossRef]
  130. Steiner, J.L.; Murphy, E.A.; McClellan, J.L.; Carmichael, M.D.; Davis, J.M. Exercise training increases mitochondrial biogenesis in the brain. J. Appl. Physiol. 2011, 111, 1066–1071. [Google Scholar] [CrossRef]
  131. Millán, I.S. The Use Of Individualized Exercise Prescription To Target Oxidative Metabolism In A Stage Iv Colorectal, Metastatic Cancer Patient: 581 May 29 1:20 PM–1:40 PM. Med. Sci. Sports Exerc. 2019, 51, 151. [Google Scholar] [CrossRef]
  132. Pickles, S.; Vigié, P.; Youle, R.J. The art of mitochondrial maintenance. Curr. Biol. CB 2018, 28, R170–R185. [Google Scholar] [CrossRef] [PubMed]
  133. Morris, G.; Walder, K.R.; Berk, M.; Marx, W.; Walker, A.J.; Maes, M.; Puri, B.K. The interplay between oxidative stress and bioenergetic failure in neuropsychiatric illnesses: Can we explain it and can we treat it? Mol. Biol. Rep. 2020, 47, 5587–5620. [Google Scholar] [CrossRef]
  134. Corona, J.C.; Duchen, M.R. PPARγ as a therapeutic target to rescue mitochondrial function in neurological disease. Free Radic. Biol. Med. 2016, 100, 153–163. [Google Scholar] [CrossRef] [PubMed]
  135. Herzig, S.; Shaw, R.J. AMPK: Guardian of metabolism and mitochondrial homeostasis. Nat. Rev. Mol. Cell Biol. 2018, 19, 121–135. [Google Scholar] [CrossRef] [PubMed]
  136. Gureev, A.P.; Sadovnikova, I.S.; Starkov, N.N.; Starkov, A.A.; Popov, V.N. p62-Nrf2-p62 Mitophagy Regulatory Loop as a Target for Preventive Therapy of Neurodegenerative Diseases. Brain Sci. 2020, 10, 847. [Google Scholar] [CrossRef]
  137. Koroshetz, W.J.; Jenkins, B.G.; Rosen, B.R.; Beal, M.F. Energy metabolism defects in Huntington’s disease and effects of coenzyme Q10. Ann. Neurol. 1997, 41, 160–165. [Google Scholar] [CrossRef] [PubMed]
  138. Li, H.; Yan, Z.; Zhu, J.; Yang, J.; He, J. Neuroprotective effects of resveratrol on ischemic injury mediated by improving brain energy metabolism and alleviating oxidative stress in rats. Neuropharmacology 2011, 60, 252–258. [Google Scholar] [CrossRef]
  139. Feng, J.; Wang, X.; Ye, X.; Ares, I.; Lopez-Torres, B.; Martínez, M.; Martinez-Larranaga, M.R.; Wang, X.; Anadón, A.; Martinez, M.A. Mitochondria as an important target of metformin: The mechanism of action, toxic and side effects, and new therapeutic applications. Pharmacol. Res. 2022, 177, 106114. [Google Scholar] [CrossRef]
  140. Rudic, J.S.; Poropat, G.; Krstic, M.N.; Bjelakovic, G.; Gluud, C. Bezafibrate for primary biliary cirrhosis. Cochrane Database Syst. Rev. 2012, 1, CD009145. [Google Scholar] [CrossRef]
  141. Jakob, T.; Nordmann, A.J.; Schandelmaier, S.; Ferreira-González, I.; Briel, M. Fibrates for primary prevention of cardiovascular disease events. Cochrane Database Syst. Rev. 2016, 2017, CD009753. [Google Scholar] [CrossRef]
  142. Douiev, L.; Sheffer, R.; Horvath, G.; Saada, A. Bezafibrate Improves Mitochondrial Fission and Function in DNM1L-Deficient Patient Cells. Cells 2020, 9, 301. [Google Scholar] [CrossRef] [PubMed]
  143. Kim, J.-Y.; Zhou, D.; Cui, X.-S. Bezafibrate prevents aging in in vitro-matured porcine oocytes. J. Anim. Sci. Technol. 2021, 63, 766–777. [Google Scholar] [CrossRef] [PubMed]
  144. Liang, S.; Yang, Z.; Hua, L.; Chen, Y.; Zhou, Y.; Ou, Y.; Chen, X.; Yue, H.; Yang, X.; Wu, X.; et al. Ciclopirox inhibits NLRP3 inflammasome activation via protecting mitochondria and ameliorates imiquimod-induced psoriatic inflammation in mice. Eur. J. Pharmacol. 2022, 930, 175156. [Google Scholar] [CrossRef] [PubMed]
  145. Minden, M.D.; Hogge, D.E.; Weir, S.J.; Kasper, J.; Webster, D.A.; Patton, L.; Jitkova, Y.; Hurren, R.; Gronda, M.; Goard, C.A.; et al. Oral ciclopirox olamine displays biological activity in a phase I study in patients with advanced hematologic malignancies. Am. J. Hematol. 2014, 89, 363–368. [Google Scholar] [CrossRef]
  146. Wang, W.-W.; Han, R.; He, H.-J.; Wang, Z.; Luan, X.-Q.; Li, J.; Feng, L.; Chen, S.-Y.; Aman, Y.; Xie, C.-L. Delineating the Role of Mitophagy Inducers for Alzheimer Disease Patients. Aging Dis. 2021, 12, 852–867. [Google Scholar] [CrossRef]
  147. Feigin, A.; Kieburtz, K.; Como, P.; Hickey, C.; Abwender, D.; Zimmerman, C.; Steinberg, K.; Shoulson, I. Assessment of coenzyme q10 tolerability in huntington’s disease. Mov. Disord. 1996, 11, 321–323. [Google Scholar] [CrossRef] [PubMed]
  148. Shults, C.W.; Haas, R.H.; Passov, D.; Beal, M.F. Coenzyme Q10 levels correlate with the activities of complexes I and II/III in mitochondria from parkinsonian and nonparkinsonian subjects. Ann. Neurol. 1997, 42, 261–264. [Google Scholar] [CrossRef] [PubMed]
  149. Ma, D.; Stokes, K.; Mahngar, K.; Domazet-Damjanov, D.; Sikorska, M.; Pandey, S. Inhibition of stress induced premature senescence in presenilin-1 mutated cells with water soluble Coenzyme Q10. Mitochondrion 2014, 17, 106–115. [Google Scholar] [CrossRef]
  150. Devos, D.; Moreau, C.; Devedjian, J.C.; Kluza, J.; Petrault, M.; Laloux, C.; Jonneaux, A.; Ryckewaert, G.; Garçon, G.; Rouaix, N.; et al. Targeting Chelatable Iron as a Therapeutic Modality in Parkinson’s Disease. Antioxid. Redox Signal. 2014, 21, 195–210. [Google Scholar] [CrossRef]
  151. Bollig, C.; Schell, L.K.; Rücker, G.; Allert, R.; Motschall, E.; Niemeyer, C.M.; Bassler, D.; Meerpohl, J.J. Deferasirox for managing iron overload in people with thalassaemia. Cochrane Database Syst. Rev. 2017, 2017, CD007476. [Google Scholar] [CrossRef]
  152. Meerpohl, J.J.; Schell, L.K.; Rücker, G.; Motschall, E.; Fleeman, N.; Niemeyer, C.M.; Bassler, D. Deferasirox for managing transfusional iron overload in people with sickle cell disease. Cochrane Database Syst. Rev. 2014, 2017, CD007477. [Google Scholar] [CrossRef] [PubMed]
  153. Pandolfo, M.; Arpa, J.; Delatycki, M.B.; Le Quan Sang, K.H.; Mariotti, C.; Munnich, A.; Sanz-Gallego, I.; Tai, G.; Tarnopolsky, M.A.; Taroni, F.; et al. Deferiprone in Friedreich Ataxia: A 6-Month Randomized Controlled Trial. Ann. Neurol. 2014, 76, 509–521. [Google Scholar] [CrossRef] [PubMed]
  154. Deferiprone to Delay Dementia (The 3D Study)—Full Text View—ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT03234686 (accessed on 29 June 2023).
  155. Saxena, D.; Spino, M.; Tricta, F.; Connelly, J.; Cracchiolo, B.M.; Hanauske, A.R.; D’Alliessi Gandolfi, D.; Mathews, M.B.; Karn, J.; Holland, B.; et al. Drug-Based Lead Discovery: The Novel Ablative Antiretroviral Profile of Deferiprone in HIV-1-Infected Cells and in HIV-Infected Treatment-Naive Subjects of a Double-Blind, Placebo-Controlled, Randomized Exploratory Trial. PLoS ONE 2016, 11, e0154842. [Google Scholar] [CrossRef]
  156. Yin, X.; Manczak, M.; Reddy, P.H. Mitochondria-targeted molecules MitoQ and SS31 reduce mutant huntingtin-induced mitochondrial toxicity and synaptic damage in Huntington’s disease. Hum. Mol. Genet. 2016, 25, 1739–1753. [Google Scholar] [CrossRef] [PubMed]
  157. Calkins, M.J.; Manczak, M.; Mao, P.; Shirendeb, U.; Reddy, P.H. Impaired mitochondrial biogenesis, defective axonal transport of mitochondria, abnormal mitochondrial dynamics and synaptic degeneration in a mouse model of Alzheimer’s disease. Hum. Mol. Genet. 2011, 20, 4515–4529. [Google Scholar] [CrossRef] [PubMed]
  158. Daubert, M.A.; Yow, E.; Dunn, G.; Marchev, S.; Barnhart, H.; Douglas, P.S.; O’Connor, C.; Goldstein, S.; Udelson, J.E.; Sabbah, H.N. Novel Mitochondria-Targeting Peptide in Heart Failure Treatment: A Randomized, Placebo-Controlled Trial of Elamipretide. Circ. Heart Fail. 2017, 10, e004389. [Google Scholar] [CrossRef]
  159. Hortmann, M.; Robinson, S.; Mohr, M.; Mauler, M.; Stallmann, D.; Reinöhl, J.; Duerschmied, D.; Peter, K.; Carr, J.; Gibson, C.M.; et al. The mitochondria-targeting peptide elamipretide diminishes circulating HtrA2 in ST-segment elevation myocardial infarction. Eur. Heart J. Acute Cardiovasc. Care 2017, 8, 695–702. [Google Scholar] [CrossRef]
  160. Mettu, P.S.; Allingham, M.J.; Cousins, S.W. Phase 1 Clinical Trial of Elamipretide in Dry Age-Related Macular Degeneration and Noncentral Geographic Atrophy. Ophthalmol. Sci. 2022, 2, 100086. [Google Scholar] [CrossRef]
  161. Karaa, A.; Bertini, E.; Carelli, V.; Cohen, B.H.; Enns, G.M.; Falk, M.J.; Goldstein, A.; Gorman, G.S.; Haas, R.; Hirano, M.; et al. Efficacy and Safety of Elamipretide in Individuals With Primary Mitochondrial Myopathy: The MMPOWER-3 Randomized Clinical Trial. Neurology 2023, 101, e238–e252. [Google Scholar] [CrossRef]
  162. Farsad-Naeimi, A.; Alizadeh, M.; Esfahani, A.; Aminabad, E.D. Effect of fisetin supplementation on inflammatory factors and matrix metalloproteinase enzymes in colorectal cancer patients. Food Funct. 2018, 9, 2025–2031. [Google Scholar] [CrossRef]
  163. Jia, S.; Xu, X.; Zhou, S.; Chen, Y.; Ding, G.; Cao, L. Fisetin induces autophagy in pancreatic cancer cells via endoplasmic reticulum stress- and mitochondrial stress-dependent pathways. Cell Death Dis. 2019, 10, 142. [Google Scholar] [CrossRef] [PubMed]
  164. Prem, P.N.; Kurian, G.A. Fisetin attenuates renal ischemia/reperfusion injury by improving mitochondrial quality, reducing apoptosis and oxidative stress. Naunyn. Schmiedebergs Arch. Pharmacol. 2022, 395, 547–561. [Google Scholar] [CrossRef] [PubMed]
  165. Shanmugam, K.; Ravindran, S.; Kurian, G.A.; Rajesh, M. Fisetin Confers Cardioprotection against Myocardial Ischemia Reperfusion Injury by Suppressing Mitochondrial Oxidative Stress and Mitochondrial Dysfunction and Inhibiting Glycogen Synthase Kinase 3β Activity. Oxidative Med. Cell. Longev. 2018, 2018, 9173436. [Google Scholar] [CrossRef] [PubMed]
  166. Verdoorn, B.P.; Evans, T.K.; Hanson, G.J.; Zhu, Y.; Prata, L.G.P.L.; Pignolo, R.J.; Atkinson, E.J.; Wissler-Gerdes, E.O.; Kuchel, G.A.; Mannick, J.B.; et al. Fisetin for COVID-19 in skilled nursing facilities: Senolytic trials in the COVID era. J. Am. Geriatr. Soc. 2021, 69, 3023–3033. [Google Scholar] [CrossRef] [PubMed]
  167. Wang, L.; Cao, D.; Wu, H.; Jia, H.; Yang, C.; Zhang, L. Fisetin Prolongs Therapy Window of Brain Ischemic Stroke Using Tissue Plasminogen Activator: A Double-Blind Randomized Placebo-Controlled Clinical Trial. Clin. Appl. Thromb. 2019, 25, 1076029619871359. [Google Scholar] [CrossRef] [PubMed]
  168. Bader, T.; Fazili, J.; Madhoun, M.; Aston, C.; Hughes, D.; Rizvi, S.; Seres, K.; Hasan, M. Fluvastatin Inhibits Hepatitis C Replication in Humans. Am. J. Gastroenterol. 2008, 103, 1383–1389. [Google Scholar] [CrossRef]
  169. Buemi, M.; Allegra, A.; Corica, F.; Aloisi, C.; Giacobbe, M.; Pettinato, G.; Corsonello, A.; Senatore, M.; Frisina, N. Effect of fluvastatin on proteinuria in patients with immunoglobulin A nephropathy. Clin. Pharmacol. Ther. 2000, 67, 427–431. [Google Scholar] [CrossRef]
  170. Paez, P.A.; Kolawole, M.; Taruselli, M.T.; Ajith, S.; Dailey, J.M.; Kee, S.A.; Haque, T.T.; Barnstein, B.O.; McLeod, J.J.A.; Caslin, H.L.; et al. Fluvastatin Induces Apoptosis in Primary and Transformed Mast Cells. Experiment 2020, 374, 104–112. [Google Scholar] [CrossRef]
  171. Ruiz-Limon, P.; Barbarroja, N.; Perez-Sanchez, C.; Aguirre, M.A.; Bertolaccini, M.L.; Khamashta, M.A.; Rodriguez-Ariza, A.; Almadén, Y.; Segui, P.; Khraiwesh, H.; et al. Atherosclerosis and cardiovascular disease in systemic lupus erythematosus: Effects of in vivo statin treatment. Ann. Rheum. Dis. 2014, 74, 1450–1458. [Google Scholar] [CrossRef]
  172. Winkler, K.; Ablethauser, C.; Gimpelewicz, C.; Bortolini, M.; Isaacsohn, J. Risk reduction and tolerability of fluvastatin in patients with the metabolic syndrome: A pooled analysis of thirty clinical trials. Clin. Ther. 2007, 29, 1987–2000. [Google Scholar] [CrossRef]
  173. Yin, Y.; Zhang, L.; Marshall, I.; Wolfe, C.; Wang, Y. Statin Therapy for Preventing Recurrent Stroke in Patients with Ischemic Stroke: A Systematic Review and Meta-Analysis of Randomized Controlled Trials and Observational Cohort Studies. Neuroepidemiology 2022, 56, 240–249. [Google Scholar] [CrossRef] [PubMed]
  174. Hu, J.; Man, W.; Shen, M.; Zhang, M.; Lin, J.; Wang, T.; Duan, Y.; Li, C.; Zhang, R.; Gao, E.; et al. Luteolin alleviates post-infarction cardiac dysfunction by up-regulating autophagy through Mst1 inhibition. J. Cell. Mol. Med. 2016, 20, 147–156. [Google Scholar] [CrossRef] [PubMed]
  175. Taliou, A.; Zintzaras, E.; Lykouras, L.; Francis, K. An Open-Label Pilot Study of a Formulation Containing the Anti-Inflammatory Flavonoid Luteolin and Its Effects on Behavior in Children With Autism Spectrum Disorders. Clin. Ther. 2013, 35, 592–602. [Google Scholar] [CrossRef] [PubMed]
  176. Versace, V.; Ortelli, P.; Dezi, S.; Ferrazzoli, D.; Alibardi, A.; Bonini, I.; Engl, M.; Maestri, R.; Assogna, M.; Ajello, V.; et al. Co-ultramicronized palmitoylethanolamide/luteolin normalizes GABAB-ergic activity and cortical plasticity in long COVID-19 syndrome. Clin. Neurophysiol. 2023, 145, 81–88. [Google Scholar] [CrossRef]
  177. Zhang, X.; Li, M.; Yue, Y.; Zhang, Y.; Wu, A. Luteoloside Prevents Sevoflurane-induced Cognitive Dysfunction in Aged Rats via Maintaining Mitochondrial Function and Dynamics in Hippocampal Neurons. Neuroscience 2023, 516, 42–53. [Google Scholar] [CrossRef]
  178. Calkin, C.V.; Chengappa, K.R.; Cairns, K.; Cookey, J.; Gannon, J.; Alda, M.; O’Donovan, C.; Reardon, C.; Sanches, M.; Růzicková, M. Treating Insulin Resistance With Metformin as a Strategy to Improve Clinical Outcomes in Treatment-Resistant Bipolar Depression (the TRIO-BD Study): A Randomized, Quadruple-Masked, Placebo-Controlled Clinical Trial. J. Clin. Psychiatry 2022, 83, 21m14022. [Google Scholar] [CrossRef]
  179. Hebrani, P.; Manteghi, A.A.; Behdani, F.; Hessami, E.; Rezayat, K.A.; Marvast, M.N.; Rezayat, A.A. Double-blind, randomized, clinical trial of metformin as add-on treatment with clozapine in treatment of schizophrenia disorder. J. Res. Med. Sci. Off. J. Isfahan Univ. Med. Sci. 2015, 20, 364–371. [Google Scholar]
  180. Ma, Z.; Liu, Z.; Li, X.; Zhang, H.; Han, D.; Xiong, W.; Zhou, H.; Yang, X.; Zeng, Q.; Ren, H.; et al. Metformin Collaborates with PINK1/Mfn2 Overexpression to Prevent Cardiac Injury by Improving Mitochondrial Function. Biology 2023, 12, 582. [Google Scholar] [CrossRef]
  181. Rizvi, F.; Sheikh, A.; Ahmed, H.; Fahmi, S.; Ikramuddin, Z.; Asif, M. Miracle medicine for prevention of migraine attack: Metformin. Prof. Med. J. 2020, 27, 812–819. [Google Scholar] [CrossRef]
  182. Zhang, K.; Wang, T.; Sun, G.-F.; Xiao, J.-X.; Jiang, L.-P.; Tou, F.-F.; Qu, X.-H.; Han, X.-J. Metformin protects against retinal ischemia/reperfusion injury through AMPK-mediated mitochondrial fusion. Free Radic. Biol. Med. 2023, 205, 47–61. [Google Scholar] [CrossRef]
  183. Zhao, M.; Li, X.W.; Chen, D.Z.; Hao, F.; Tao, S.X.; Yu, H.Y.; Cheng, R.; Liu, H. Neuro-Protective Role of Metformin in Patients with Acute Stroke and Type 2 Diabetes Mellitus via AMPK/Mammalian Target of Rapamycin (mTOR) Signaling Pathway and Oxidative Stress. Experiment 2019, 25, 2186–2194. [Google Scholar] [CrossRef] [PubMed]
  184. Deng, Y.; Wang, R.; Li, S.; Zhu, X.; Wang, T.; Wu, J.; Zhang, J. Methylene blue reduces incidence of early postoperative cognitive disorders in elderly patients undergoing major non-cardiac surgery: An open–label randomized controlled clinical trial. J. Clin. Anesthesia 2020, 68, 110108. [Google Scholar] [CrossRef]
  185. Singh, N.; MacNicol, E.; DiPasquale, O.; Randall, K.; Lythgoe, D.; Mazibuko, N.; Simmons, C.; Selvaggi, P.; Stephenson, S.; Turkheimer, F.E.; et al. The effects of acute Methylene Blue administration on cerebral blood flow and metabolism in humans and rats. J. Cereb. Blood Flow Metab. 2023. OnlineFirst. [Google Scholar] [CrossRef] [PubMed]
  186. Shiells, H.; Schelter, B.O.; Bentham, P.; Baddeley, T.C.; Rubino, C.M.; Ganesan, H.; Hammel, J.; Vuksanovic, V.; Staff, R.T.; Murray, A.D.; et al. Concentration-Dependent Activity of Hydromethylthionine on Clinical Decline and Brain Atrophy in a Randomized Controlled Trial in Behavioral Variant Frontotemporal Dementia. J. Alzheimer’s Dis. 2020, 75, 501–519. [Google Scholar] [CrossRef] [PubMed]
  187. Zhao, N.; Liang, P.; Zhuo, X.; Su, C.; Zong, X.; Guo, B.; Han, D.; Yan, X.; Hu, S.; Zhang, Q.; et al. After Treatment with Methylene Blue is Effective against Delayed Encephalopathy after Acute Carbon Monoxide Poisoning. Basic Clin. Pharmacol. Toxicol. 2017, 122, 470–480. [Google Scholar] [CrossRef] [PubMed]
  188. Zoellner, L.A.; Telch, M.; Foa, E.B.; Farach, F.J.; McLean, C.P.; Gallop, R.; Bluett, E.J.; Cobb, A.; Gonzalez-Lima, F. Enhancing Extinction Learning in Posttraumatic Stress Disorder With Brief Daily Imaginal Exposure and Methylene Blue: A Randomized Controlled Trial. J. Clin. Psychiatry 2017, 78, e782–e789. [Google Scholar] [CrossRef]
  189. Sadovnikova, I.S.; Gureev, A.P.; Ignatyeva, D.A.; Gryaznova, M.V.; Chernyshova, E.V.; Krutskikh, E.P.; Novikova, A.G.; Popov, V.N. Nrf2/ARE Activators Improve Memory in Aged Mice via Maintaining of Mitochondrial Quality Control of Brain and the Modulation of Gut Microbiome. Pharmaceuticals 2021, 14, 607. [Google Scholar] [CrossRef]
  190. Wischik, C.M.; Staff, R.T.; Wischik, D.J.; Bentham, P.; Murray, A.D.; Storey, J.M.; Kook, K.A.; Harrington, C.R. Tau Aggregation Inhibitor Therapy: An Exploratory Phase 2 Study in Mild or Moderate Alzheimer’s Disease. J. Alzheimers Dis. 2015, 44, 705–720. [Google Scholar] [CrossRef]
  191. Kshirsagar, S.; Sawant, N.; Morton, H.; Reddy, A.P.; Reddy, P.H. Protective effects of mitophagy enhancers against amyloid beta-induced mitochondrial and synaptic toxicities in Alzheimer disease. Hum. Mol. Genet. 2021, 31, 423–439. [Google Scholar] [CrossRef]
  192. Chen, A.C.; Martin, A.J.; Choy, B.; Fernández-Peñas, P.; Dalziell, R.A.; McKenzie, C.A.; Scolyer, R.A.; Dhillon, H.M.; Vardy, J.L.; Kricker, A.; et al. A Phase 3 Randomized Trial of Nicotinamide for Skin-Cancer Chemoprevention. N. Engl. J. Med. 2015, 373, 1618–1626. [Google Scholar] [CrossRef]
  193. Gale, E.A.M.; Bingley, P.J.; Emmett, C.L.; Collier, T.; European Nicotinamide Diabetes Intervention Trial (ENDIT) Group. European Nicotinamide Diabetes Intervention Trial (ENDIT): A randomised controlled trial of intervention before the onset of type 1 diabetes. Lancet 2004, 363, 925–931. [Google Scholar] [CrossRef] [PubMed]
  194. Libri, V.; Yandim, C.; Athanasopoulos, S.; Loyse, N.; Natisvili, T.; Law, P.P.; Chan, P.K.; Mohammad, T.; Mauri, M.; Tam, K.T.; et al. Epigenetic and neurological effects and safety of high-dose nicotinamide in patients with Friedreich’s ataxia: An exploratory, open-label, dose-escalation study. Lancet 2014, 384, 504–513. [Google Scholar] [CrossRef] [PubMed]
  195. Chong, R.; Wakade, C.; Seamon, M.; Giri, B.; Morgan, J.; Purohit, S. Niacin Enhancement for Parkinson’s Disease: An Effectiveness Trial. Front. Aging Neurosci. 2021, 13, 667032. [Google Scholar] [CrossRef] [PubMed]
  196. Tully, L.; Humiston, J.; Cash, A. Oxaloacetate reduces emotional symptoms in premenstrual syndrome (PMS): Results of a placebo-controlled, cross-over clinical trial. Obstet. Gynecol. Sci. 2020, 63, 195–204. [Google Scholar] [CrossRef] [PubMed]
  197. Cash, A.; Kaufman, D.L. Oxaloacetate Treatment For Mental And Physical Fatigue In Myalgic Encephalomyelitis/Chronic Fatigue Syndrome (ME/CFS) and Long-COVID fatigue patients: A non-randomized controlled clinical trial. J. Transl. Med. 2022, 20, 295. [Google Scholar] [CrossRef]
  198. Wilkins, H.M.; Koppel, S.; Carl, S.M.; Ramanujan, S.; Weidling, I.; Michaelis, M.L.; Michaelis, E.K.; Swerdlow, R.H. Oxaloacetate enhances neuronal cell bioenergetic fluxes and infrastructure. J. Neurochem. 2016, 137, 76–87. [Google Scholar] [CrossRef] [PubMed]
  199. Swerdlow, R.H.; Lyons, K.E.; Khosla, S.K.; Nashatizadeh, M.; Pahwa, R. A Pilot Study of Oxaloacetate 100 mg Capsules in Parkinson’s Disease Patients. J. Park. Dis. Alzheimers Dis. 2016, 3, 4. [Google Scholar]
  200. Flicker, L.; Evans, J.G. Piracetam for dementia or cognitive impairment. Cochrane Database Syst. Rev. 2001, CD001011. [Google Scholar] [CrossRef]
  201. Hofmeyr, G.J.; Kulier, R. Piracetam for fetal distress in labour. Cochrane Database Syst. Rev. 2012, 2012, CD001064. [Google Scholar] [CrossRef]
  202. Ricci, S.; Celani, M.G.; Cantisani, T.A.; Righetti, E. Piracetam for acute ischaemic stroke. Cochrane Database Syst. Rev. 2012, 2012, CD000419. [Google Scholar] [CrossRef]
  203. Sivalingam, K.; Samikkannu, T. Neuroprotective Effect of Piracetam against Cocaine-Induced Neuro Epigenetic Modification of DNA Methylation in Astrocytes. Brain Sci. 2020, 10, 611. [Google Scholar] [CrossRef] [PubMed]
  204. Stockburger, C.; Kurz, C.; Koch, K.A.; Eckert, S.H.; Leuner, K.; Müller, W.E. Improvement of mitochondrial function and dynamics by the metabolic enhancer piracetam. Biochem. Soc. Trans. 2013, 41, 1331–1334. [Google Scholar] [CrossRef] [PubMed]
  205. Nagalla, S.; Ballas, S.K. Drugs for preventing red blood cell dehydration in people with sickle cell disease. Cochrane Database Syst. Rev. 2018, 10, CD003426. [Google Scholar] [CrossRef] [PubMed]
  206. Shi, Q.; Zhao, G.; Wei, S.; Guo, C.; Wu, X.; Zhao, R.C.; Di, G. Pterostilbene alleviates liver ischemia/reperfusion injury via PINK1-mediated mitophagy. J. Pharmacol. Sci. 2021, 148, 19–30. [Google Scholar] [CrossRef] [PubMed]
  207. Jensen, J.B.; Dollerup, O.L.; Møller, A.B.; Billeskov, T.B.; Dalbram, E.; Chubanava, S.; Damgaard, M.V.; Dellinger, R.W.; Trošt, K.; Moritz, T.; et al. A randomized placebo-controlled trial of nicotinamide riboside and pterostilbene supplementation in experimental muscle injury in elderly individuals. J. Clin. Investig. 2022, 7, e158314. [Google Scholar] [CrossRef] [PubMed]
  208. Simic, P.; Parada, X.F.V.; Parikh, S.M.; Dellinger, R.; Guarente, L.P.; Rhee, E.P. Nicotinamide riboside with pterostilbene (NRPT) increases NAD+ in patients with acute kidney injury (AKI): A randomized, double-blind, placebo-controlled, stepwise safety study of escalating doses of NRPT in patients with AKI. BMC Nephrol. 2020, 21, 342. [Google Scholar] [CrossRef]
  209. Steiner, A.Z.; Hansen, K.R.; Barnhart, K.T.; Cedars, M.I.; Legro, R.S.; Diamond, M.P.; Krawetz, S.A.; Usadi, R.; Baker, V.L.; Coward, R.M.; et al. The effect of antioxidants on male factor infertility: The Males, Antioxidants, and Infertility (MOXI) randomized clinical trial. Fertil. Steril. 2020, 113, 552–560.e3. [Google Scholar] [CrossRef]
  210. Zheng, J.; Liu, W.; Zhu, X.; Ran, L.; Lang, H.; Yi, L.; Mi, M.; Zhu, J. Pterostilbene Enhances Endurance Capacity via Promoting Skeletal Muscle Adaptations to Exercise Training in Rats. Molecules 2020, 25, 186. [Google Scholar] [CrossRef]
  211. De la Rubia, J.E.; Drehmer, E.; Platero, J.L.; Benlloch, M.; Caplliure-Llopis, J.; Villaron-Casales, C.; de Bernardo, N.; AlarcÓn, J.; Fuente, C.; Carrera, S.; et al. Efficacy and tolerability of EH301 for amyotrophic lateral sclerosis: A randomized, double-blind, placebo-controlled human pilot study. Amyotroph. Lateral Scler. Front. Degener. 2019, 20, 115–122. [Google Scholar] [CrossRef]
  212. MJeyaraman, M.M.; Al-Yousif, N.S.H.; Mann, A.S.; Dolinsky, V.W.; Rabbani, R.; Zarychanski, R.; Abou-Setta, A.M. Resveratrol for adults with type 2 diabetes mellitus. Cochrane Database Syst. Rev. 2020, 2020, CD011919. [Google Scholar]
  213. McCreary, M.R.; Schnell, P.M.; Rhoda, D.A. Randomized double-blind placebo-controlled proof-of-concept trial of resveratrol for outpatient treatment of mild coronavirus disease (COVID-19). Sci. Rep. 2022, 12, 10978. [Google Scholar] [CrossRef] [PubMed]
  214. Sawda, C.; Moussa, C.; Turner, R.S. Resveratrol for Alzheimer’s disease. Ann. N. Y. Acad. Sci. 2017, 1403, 142–149. [Google Scholar] [CrossRef] [PubMed]
  215. Zheng, M.; Bai, Y.; Sun, X.; Fu, R.; Liu, L.; Liu, M.; Li, Z.; Huang, X. Resveratrol Reestablishes Mitochondrial Quality Control in Myocardial Ischemia/Reperfusion Injury through Sirt1/Sirt3-Mfn2-Parkin-PGC-1α Pathway. Molecules 2022, 27, 5545. [Google Scholar] [CrossRef] [PubMed]
  216. Zortea, K.; Franco, V.C.; Francesconi, L.P.; Cereser, K.M.M.; Lobato, M.I.R.; Belmonte-De-Abreu, P.S. Resveratrol Supplementation in Schizophrenia Patients: A Randomized Clinical Trial Evaluating Serum Glucose and Cardiovascular Risk Factors. Nutrients 2016, 8, 73. [Google Scholar] [CrossRef]
  217. Aranda-Rivera, A.K.; Cruz-Gregorio, A.; Aparicio-Trejo, O.E.; Tapia, E.; Sánchez-Lozada, L.G.; García-Arroyo, F.E.; Amador-Martínez, I.; Orozco-Ibarra, M.; Fernández-Valverde, F.; Pedraza-Chaverri, J. Sulforaphane Protects against Unilateral Ureteral Obstruction-Induced Renal Damage in Rats by Alleviating Mitochondrial and Lipid Metabolism Impairment. Antioxidants 2022, 11, 1854. [Google Scholar] [CrossRef]
  218. Ghazizadeh-Hashemi, F.; Bagheri, S.; Ashraf-Ganjouei, A.; Moradi, K.; Shahmansouri, N.; Mehrpooya, M.; Noorbala, A.A.; Akhondzadeh, S. Efficacy and safety of sulforaphane for treatment of mild to moderate depression in patients with history of cardiac interventions: A randomized, double-blind, placebo-controlled clinical trial. Psychiatry Clin. Neurosci. 2021, 75, 250–255. [Google Scholar] [CrossRef]
  219. Shiina, A.; Kanahara, N.; Sasaki, T.; Oda, Y.; Hashimoto, T.; Hasegawa, T.; Yoshida, T.; Iyo, M.; Hashimoto, K. An Open Study of Sulforaphane-rich Broccoli Sprout Extract in Patients with Schizophrenia. Clin. Psychopharmacol. Neurosci. 2015, 13, 62–67. [Google Scholar] [CrossRef]
  220. Zimmerman, A.W.; Singh, K.; Connors, S.L.; Liu, H.; Panjwani, A.A.; Lee, L.C.; Diggins, E.; Foley, A.; Melnyk, S.; Singh, I.N.; et al. Randomized controlled trial of sulforaphane and metabolite discovery in children with Autism Spectrum Disorder. Mol. Autism 2021, 12, 38. [Google Scholar] [CrossRef]
  221. Savencu, C.E.; Linţa, A.; Farcaş, G.; Bînă, A.M.; Creţu, O.M.; Maliţa, D.C.; Muntean, D.M.; Sturza, A. Impact of Dietary Restriction Regimens on Mitochondria, Heart, and Endothelial Function: A Brief Overview. Front. Physiol. 2021, 12, 768383. [Google Scholar] [CrossRef]
  222. López-Torres, M.; Gredilla, R.; Sanz, A.; Barja, G. Influence of aging and long-term caloric restriction on oxygen radical generation and oxidative DNA damage in rat liver mitochondria. Free Radic. Biol. Med. 2002, 32, 882–889. [Google Scholar] [CrossRef]
  223. Pan, J.W.; Rothman, D.L.; Behar, K.L.; Stein, D.T.; Hetherington, H.P. Human Brain β-Hydroxybutyrate and Lactate Increase in Fasting-Induced Ketosis. J. Cereb. Blood Flow Metab. 2000, 20, 1502–1507. [Google Scholar] [CrossRef]
  224. Pan, J.W.; Telang, F.W.; Lee, J.H.; De Graaf, R.A.; Rothman, D.L.; Stein, D.T.; Hetherington, H.P. Measurement of β-hydroxybutyrate in acute hyperketonemia in human brain. J. Neurochem. 2001, 79, 539–544. [Google Scholar] [CrossRef]
  225. Bough, K.J.; Wetherington, J.; Hassel, B.; Pare, J.F.; Gawryluk, J.W.; Greene, J.G.; Shaw, R.; Smith, Y.; Geiger, J.D.; Dingledine, R.J. Mitochondrial biogenesis in the anticonvulsant mechanism of the ketogenic diet. Ann. Neurol. 2006, 60, 223–235. [Google Scholar] [CrossRef] [PubMed]
  226. Elamin, M.; Ruskin, D.N.; Masino, S.A.; Sacchetti, P. Ketone-Based Metabolic Therapy: Is Increased NAD+ a Primary Mechanism? Front. Mol. Neurosci. 2017, 10, 377. [Google Scholar] [CrossRef] [PubMed]
  227. Patergnani, S.; Bouhamida, E.; Leo, S.; Pinton, P.; Rimessi, A. Mitochondrial Oxidative Stress and “Mito-Inflammation”: Actors in the Diseases. Biomedicines 2021, 9, 216. [Google Scholar] [CrossRef]
  228. Manosalva, C.; Quiroga, J.; Hidalgo, A.I.; Alarcón, P.; Anseoleaga, N.; Hidalgo, M.A.; Burgos, R.A. Role of Lactate in Inflammatory Processes: Friend or Foe. Front. Immunol. 2022, 12, 808799. [Google Scholar] [CrossRef]
  229. Tauffenberger, A.; Fiumelli, H.; Almustafa, S.; Magistretti, P.J. Lactate and pyruvate promote oxidative stress resistance through hormetic ROS signaling. Cell Death Dis. 2019, 10, 9. [Google Scholar] [CrossRef] [PubMed]
  230. Gomes, A.P.; Price, N.L.; Ling, A.J.Y.; Moslehi, J.J.; Montgomery, M.K.; Rajman, L.; White, J.P.; Teodoro, J.S.; Wrann, C.D.; Hubbard, B.P.; et al. Declining NAD+ Induces a Pseudohypoxic State Disrupting Nuclear-Mitochondrial Communication during Aging. Cell 2013, 155, 1624–1638. [Google Scholar] [CrossRef]
  231. Bosco, G.; Yang, Z.; Nandi, J.; Wang, J.; Chen, C.; Camporesi, E.M. Effects of Hyperbaric Oxygen on Glucose, Lactate, Glycerol and Anti-Oxidant Enzymes in the Skeletal Muscle of Rats During Ischaemia and Reperfusion. Clin. Exp. Pharmacol. Physiol. 2007, 34, 70–76. [Google Scholar] [CrossRef]
  232. Corbett, J.; Tipton, M.J.; Perissiou, M.; James, T.; Young, J.S.; Newman, A.; Cummings, M.; Montgomery, H.; Grocott, M.P.W.; Shepherd, A.I. Effect of different levels of acute hypoxia on subsequent oral glucose tolerance in males with overweight: A balanced cross-over pilot feasibility study. Physiol. Rep. 2023, 11, e15623. [Google Scholar] [CrossRef]
  233. Reiter, R.J.; Ma, Q.; Sharma, R. Melatonin in Mitochondria: Mitigating Clear and Present Dangers. Physiology 2020, 35, 86–95. [Google Scholar] [CrossRef] [PubMed]
  234. Powell, C.L.; Davidson, A.R.; Brown, A.M. Universal Glia to Neurone Lactate Transfer in the Nervous System: Physiological Functions and Pathological Consequences. Biosensors 2020, 10, 183. [Google Scholar] [CrossRef] [PubMed]
  235. Bonn, J.A.; Harrison, J.; Rees, W.L. Lactate-Induced Anxiety: Therapeutic Application. Br. J. Psychiatry 1971, 119, 468–470. [Google Scholar] [CrossRef] [PubMed]
  236. Auriacombe, M.; Rénéric, J.P.; Usandizaga, D.; Gomez, F.; Combourieu, I.; Tignol, J. Post-ECT agitation and plasma lactate concentrations. J. ECT 2000, 16, 263–267. [Google Scholar] [CrossRef]
  237. Trangmar, S.J.; Chiesa, S.T.; Kalsi, K.K.; Secher, N.H.; González-Alonso, J. Whole body hyperthermia, but not skin hyperthermia, accelerates brain and locomotor limb circulatory strain and impairs exercise capacity in humans. Physiol. Rep. 2017, 5, e13108. [Google Scholar] [CrossRef]
  238. Koush, Y.; de Graaf, R.A.; Jiang, L.; Rothman, D.L.; Hyder, F. Functional MRS with J-edited lactate in human motor cortex at 4 T. Neuroimage 2018, 184, 101–108. [Google Scholar] [CrossRef]
  239. Koush, Y.; Rothman, D.L.; Behar, K.L.; de Graaf, R.A.; Hyder, F. Human brain functional MRS reveals interplay of metabolites implicated in neurotransmission and neuroenergetics. J. Cereb. Blood Flow Metab. 2022, 42, 911–934. [Google Scholar] [CrossRef]
  240. Crémillieux, Y.; Dumont, U.; Mazuel, L.; Salvati, R.; Zhendre, V.; Rizzitelli, S.; Blanc, J.; Roumes, H.; Pinaud, N.; Bouzier-Sore, A.-K. Online Quantification of Lactate Concentration in Microdialysate During Cerebral Activation Using 1H-MRS and Sensitive NMR Microcoil. Front. Cell Neurosci. 2019, 13, 89. [Google Scholar] [CrossRef]
  241. Dogan, A.E.; Yuksel, C.; Du, F.; Chouinard, V.-A.; Öngür, D. Brain lactate and pH in schizophrenia and bipolar disorder: A systematic review of findings from magnetic resonance studies. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2018, 43, 1681–1690. [Google Scholar] [CrossRef]
  242. Maddock, R.J.; Buonocore, M.H. MR Spectroscopic Studies of the Brain in Psychiatric Disorders. In Brain Imaging in Behavioral Neuroscience; Carter, C.S., Dalley, J.W., Eds.; In Current Topics in Behavioral Neurosciences; Springer: Berlin/Heidelberg, Germany, 2012; pp. 199–251. [Google Scholar]
  243. Corrigan, N.M.; Richards, T.L.; Friedman, S.D.; Petropoulos, H.; Dager, S.R. Improving 1H MRSI measurement of cerebral lactate for clinical applications. Psychiatry Res. Neuroimaging 2010, 182, 40–47. [Google Scholar] [CrossRef]
  244. Maddock, R.J.; Buonocore, M.H.; Miller, A.R.; Yoon, J.H.; Soosman, S.K.; Unruh, A.M. Abnormal Activity-Dependent Brain Lactate and Glutamate+Glutamine Responses in Panic Disorder. Biol. Psychiatry 2013, 73, 1111–1119. [Google Scholar] [CrossRef] [PubMed]
  245. Karnib, N.; El-Ghandour, R.; El Hayek, L.; Nasrallah, P.; Khalifeh, M.; Barmo, N.; Jabre, V.; Ibrahim, P.; Bilen, M.; Stephan, J.S.; et al. Lactate is an antidepressant that mediates resilience to stress by modulating the hippocampal levels and activity of histone deacetylases. Neuropsychopharmacology 2019, 44, 1152–1162. [Google Scholar] [CrossRef] [PubMed]
  246. Hashimoto, T.; Tsukamoto, H.; Takenaka, S.; Olesen, N.D.; Petersen, L.G.; Sørensen, H.; Nielsen, H.B.; Secher, N.H.; Ogoh, S. Maintained exercise-enhanced brain executive function related to cerebral lactate metabolism in men. FASEB J. 2018, 32, 1417–1427. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The implication of altered metabolic health on the molecular, intracellular, tissue level, and neurobehavioural effects of lactate in the brain. Metabolic syndrome (MS) is associated with mitochondrial dysfunction and impaired lactate oxidation. (A) In response to incremental exercise testing, participants with MS generate more circulating lactate at lower power outputs compared to individuals who do not have MS. Trained athletes who are metabolically fit can produce relatively less lactate at higher outputs. (B) In response to visual stimulus, participants with panic disorder produce more lactate in the visual cortex compared to a healthy control group. (C) In response to a personalized exercise prescription applied over 12 months, pre-type 2 diabetes was reversed, and the lactate response to incremental physical exercise was improved to reflect an improvement in mitochondrial function. Elevated concentrations of circulating or brain lactate can have effects on the brain, signalling at the molecular, cellular, tissue, and brain nuclei level to influence behaviour which map onto Research Domain Criteria (RDoC) domains. Figure inspired by [65], with graphical data adapted from [3,35,88].
Figure 1. The implication of altered metabolic health on the molecular, intracellular, tissue level, and neurobehavioural effects of lactate in the brain. Metabolic syndrome (MS) is associated with mitochondrial dysfunction and impaired lactate oxidation. (A) In response to incremental exercise testing, participants with MS generate more circulating lactate at lower power outputs compared to individuals who do not have MS. Trained athletes who are metabolically fit can produce relatively less lactate at higher outputs. (B) In response to visual stimulus, participants with panic disorder produce more lactate in the visual cortex compared to a healthy control group. (C) In response to a personalized exercise prescription applied over 12 months, pre-type 2 diabetes was reversed, and the lactate response to incremental physical exercise was improved to reflect an improvement in mitochondrial function. Elevated concentrations of circulating or brain lactate can have effects on the brain, signalling at the molecular, cellular, tissue, and brain nuclei level to influence behaviour which map onto Research Domain Criteria (RDoC) domains. Figure inspired by [65], with graphical data adapted from [3,35,88].
Antioxidants 12 01656 g001
Figure 2. Therapeutic interventions that could influence lactate metabolism in the brain. Lactate concentration can be manipulated using metabolic therapies and subsequent changes measured with functional magnetic resonance spectroscopy (fMRS) in the brain or serum lactate in peripheral circulation (see dotted arrows). The concentration of lactate in the brain is primarily dependent on the rate of production influenced by the rates of glycolysis and glycogenolysis in the brain and body. On the other hand, the rate of lactate consumption relies on brain mitochondria. Mitochondria could be therapeutically manipulated in various ways (see solid arrows), and the outcome of interventions could be measured by observing changes in lactate dynamics.
Figure 2. Therapeutic interventions that could influence lactate metabolism in the brain. Lactate concentration can be manipulated using metabolic therapies and subsequent changes measured with functional magnetic resonance spectroscopy (fMRS) in the brain or serum lactate in peripheral circulation (see dotted arrows). The concentration of lactate in the brain is primarily dependent on the rate of production influenced by the rates of glycolysis and glycogenolysis in the brain and body. On the other hand, the rate of lactate consumption relies on brain mitochondria. Mitochondria could be therapeutically manipulated in various ways (see solid arrows), and the outcome of interventions could be measured by observing changes in lactate dynamics.
Antioxidants 12 01656 g002
Table 1. Compounds that target mitochondrial function known to penetrate the CNS.
Table 1. Compounds that target mitochondrial function known to penetrate the CNS.
CompoundTargetMechanismClinical TrialsReferences
BezafibratePPARα↓Fission
↑Mitophagy
↑Biogenesis
Primary Biliary Cirrhosis
Cardiovascular Disease
[140,141,142,143]
Ciclopirox olaminePGC-1α
Metal Chelator
↑Mitophagy
↑Biogenesis
Anti-tumour[144,145,146]
Coenzyme Q10Electron Acceptor↑MitophagyHuntington’s Disease
Parkinson’s Disease
[147,148,149]
DeferiproneIron Chelator↑MitophagyParkinson’s Disease
Friedrich’s Ataxia
Alzheimer’s Disease
HIV
Thalassaemia
Sickle Cell Disease
[150,151,152,153,154,155]
ElamipretideCardiolipin
Stabiliser
↓Fission
↑Fusion
↑Mitophagy
↑Biogenesis
1° Mitochondrial Myopathy
Macular Degeneration
Myocardial Infarction
Heart Failure
[156,157,158,159,160,161]
FisetinAntioxidant
SIRT1/AMPK
PPARγ
↓Fission
↑Fusion
↑Mitophagy
↑Biogenesis
COVID-19
Cancer
Stroke
[162,163,164,165,166,167]
FluvastatinHMG-CoA Reductase↑Mitophagy
↑Biogenesis
Systemic lupus erythematosus
Metabolic syndrome
Hepatitis C
[168,169,170,171,172,173]
LuteolinFlavanoid
Multiple Targets
↓Fission
↑Fusion
↑Mitophagy
↑Biogenesis
COVID-1916/08/2023 18:44:00
Autism
[174,175,176,177]
MetforminAMPK↓Fission
↑Fusion
↑Mitophagy
↑Biogenesis
Schizophrenia
Bipolar Affective Disorder
Migraine
Stroke
[178,179,180,181,182,183]
Methylene BlueNRF2
Electron Recycling
↓Fission
↑Fusion
↑Mitophagy
↑Biogenesis
Post-traumatic Stress Disorder
Bipolar Affective Disorder
Alzheimer’s Disease
Post-operative Delirium
Frontotemporal Dementia
[184,185,186,187,188,189,190]
NicotinamideSIRT1↓Fission
↑Fusion
↑Mitophagy
↑Biogenesis
Friedrich’s Ataxia
Skin Cancer
Diabetes
Parkinson’s Disease
[191,192,193,194,195]
OxaloacetateAntioxidant
TCA cycle
SIRT1
↑BiogenesisCOVID-19
Chronic Fatigue Syndrome
Parkinson’s Disease
Premenstrual Syndrome
[196,197,198,199]
PiracetamNRF2↓Fission
↑Fusion
↑Biogenesis
Stroke
Dementia
Sickle Cell Disease
Foetal Distress in Labour
[200,201,202,203,204,205]
PterostilbenePI3K-Akt-mTOR
SIRT1
↑Mitophagy
↑Biogenesis
Infertility
Muscle Injury
Amyotrophic Lateral Sclerosis
Acute Kidney Injury
[206,207,208,209,210,211]
ResveratrolPPARs↓Fission
↓↑Fusion
↑Mitophagy
↑Biogenesis
Diabetes
Schizophrenia
Alzheimer’s Disease
COVID-19
[212,213,214,215,216]
SulforaphaneNRF2↓Fission
↑Fusion
↑Mitophagy
↑Biogenesis
Autism
Depression
Schizophrenia
[217,218,219,220]
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

Caddye, E.; Pineau, J.; Reyniers, J.; Ronen, I.; Colasanti, A. Lactate: A Theranostic Biomarker for Metabolic Psychiatry? Antioxidants 2023, 12, 1656. https://doi.org/10.3390/antiox12091656

AMA Style

Caddye E, Pineau J, Reyniers J, Ronen I, Colasanti A. Lactate: A Theranostic Biomarker for Metabolic Psychiatry? Antioxidants. 2023; 12(9):1656. https://doi.org/10.3390/antiox12091656

Chicago/Turabian Style

Caddye, Edward, Julien Pineau, Joshua Reyniers, Itamar Ronen, and Alessandro Colasanti. 2023. "Lactate: A Theranostic Biomarker for Metabolic Psychiatry?" Antioxidants 12, no. 9: 1656. https://doi.org/10.3390/antiox12091656

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