Hostname: page-component-76fb5796d-qxdb6 Total loading time: 0 Render date: 2024-04-25T08:01:41.191Z Has data issue: false hasContentIssue false

Glutamine metabolism and optimal immune and CNS function

Published online by Cambridge University Press:  26 October 2022

Philip Newsholme*
Affiliation:
Curtin Medical School, Faculty of Health Sciences and Curtin Health Innovation Research Institute, Curtin University, Perth, Western Australia, Australia
Vinicius Leonardo Sousa Diniz
Affiliation:
Curtin Medical School, Faculty of Health Sciences and Curtin Health Innovation Research Institute, Curtin University, Perth, Western Australia, Australia Interdisciplinary Post-Graduate Program in Health Sciences, Cruzeiro do Sul University, São Paulo, SP, Brazil
Garron Thomas Dodd
Affiliation:
Metabolic Neuroscience Laboratory, Department of Anatomy and Physiology, The University of Melbourne, Melbourne, VIC 3010, Australia
Vinicius Cruzat
Affiliation:
Torrens University Australia, Brisbane, Queensland, Australia Faculty of Health, Southern Cross University, Gold Coast, Queensland, Australia
*
*Corresponding author: Philip Newsholme, email philip.newsholme@curtin.edu.au
Rights & Permissions [Opens in a new window]

Abstract

Nutrients can impact and regulate cellular metabolism and cell function which is particularly important for the activation and function of diverse immune subsets. Among the critical nutrients for immune cell function and fate, glutamine is possibly the most widely recognised immunonutrient, playing key roles in TCA cycle, heat shock protein responses and antioxidant systems. In addition, glutamine is also involved with inter-organ ammonia transport, and this is particularly important for not only immune cells, but also to the brain, especially in catabolic situations such as critical care and extenuating exercise. The well characterised fall in blood glutamine availability has been the main reason for studies to investigate the possible effects of glutamine replacement via supplementation but many of the results are in poor agreement. At the same time, a range of complex pathways involved in glutamine metabolism have been revealed via supplementation studies. This article will briefly review the function of glutamine in the immune system, with emphasis on metabolic mechanisms, and the emerging role of glutamine in the brain glutamate/gamma-amino butyric acid cycle. In addition, relevant aspects of glutamine supplementation are discussed.

Type
Conference on ‘Nutrition, immune function and infectious disease’
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press on behalf of The Nutrition Society

Nutritional immunology describes the influence of nutrition on the immune system, anti-viral activity and associated protective functions(Reference Wu, Lewis and Pae1,Reference Calder, Carr and Gombart2) . Deficiency in macronutrients and/or micronutrients causes impairment of immune function, which can be reversed by nutrient repletion. Among the nutrients that have important immunological roles and/or recovery from injuries/trauma, glutamine is probably the most widely recognised immunonutrient, as it is required for key immune-inflammatory responses. Glutamine plays various essential roles in almost every cell in the body, involving complex and dynamic regulations.

Although several amino acids play important roles in cell metabolism, glutamine is key for the intermediary metabolism and inter-organ nitrogen exchange via ammonia (NH3) transport between tissues. Glutamine can also be oxidised in the TCA cycle as an energy source, serve as a substrate for nucleotide synthesis, NAD+ and NADPH synthesis(Reference Curi, Newsholme and Marzuca-Nassr3,Reference Cruzat, Macedo Rogero and Noel Keane4) . In addition, glutamine has antioxidant properties by acting as a primary glutamate donor to the synthesis of γ-l-glutamyl-l-cysteinyl-glycine (also known as glutathione – GSH), and also chaperone properties by modulating the expression of heat shock proteins (HSPs)(Reference Cruzat, Macedo Rogero and Noel Keane4).

Glutamine is the most abundant amino acid in the blood, achieving concentrations of 2× to 100× greater than other amino acids with a concentration range of approximately 0⋅5–0⋅7 mm. Normal levels of glutamine in the body are usually achieved by a balanced diet associated with appropriate physical activity. However, it is already established that in catabolic situations, such as burns, post-trauma/surgery and exhaustive exercise, glutamine exogenous supply might be required due to the inability of the body to maintain optimal levels of the amino acid under some conditions (Fig. 1)(Reference Smedberg, Rooyackers and Norberg5). Glutamine may be required to attenuate immunodepression and help the body to recover more efficiently from an excess of damage and inflammation. Although glutamine exogenous supply seems a very plausible solution, it has always been a matter of vigorous debate. In this review, we aim to discuss the traditional and conceptual roles of glutamine, as well as its antioxidant and chaperone properties. We also describe an emergent glutamine role in brain physiology via the glutamate/gamma-amino butyric acid (GABA) cycle (GGC) in the context of exercise, and finally conclude with current considerations for translating possible glutamine supplementation effects for catabolic situations.

Fig. 1. Blood glutamine concentration changes according to the balance between major organ producers and consumers in health and catabolic situations. As the most abundant and versatile amino acid in the body, glutamine largely depends on the balance between its synthesis, release and uptake by organs and tissues. In turn, multiple intracellular pathways require glutamine as a substrate to maintain homoeostasis. In health, there is a balance between glutamine synthesis and degradation, while in catabolic situations organs responsible for glutamine synthesis reduce its production, such as the skeletal muscle tissue. Other glutamine producers, such as the adipose tissue and the lungs, do not have the capability to replenish the needs of the amino acid during catabolism. Moreover, the liver, a main glutamine producer in health becomes a major glutamine consumer under disease conditions, due to gluconeogenesis support. At the same time, cells of the immune system increase their demand for both glutamine and glucose. Although the brain and kidneys may have their glutamine capabilities altered according to the type of disease/catabolic condition, no significant changes may counteract the fall in glutamine (GLN) availability.

Brief overview of glutamine biochemistry and metabolism

Two principal enzymes regulate intracellular glutamine metabolism. Glutamine synthetase (GS; EC 6⋅3⋅1⋅2) catalyses the synthesis of glutamine from glutamate and ammonia in glutamine synthesising cells and tissues, while phosphate-dependent glutaminase (GLS; EC 3⋅5⋅1⋅2) catalyses the hydrolysis of glutamine to glutamate (with the release of ammonia). Glutamine availability largely depends on the balance between its synthesis, release and uptake by organs and tissues. The lungs, liver, brain, adipose tissue and skeletal muscles are the major glutamine suppliers as they exhibit high activity of tissue-specific glutamine synthesis, i.e. GS. In contrast, through the activity of GLS and other enzymes capable of degrading glutamine, the amino acid can undergo various rates of degradation. Cells of the intestinal mucosa, renal tubule and especially leucocytes are the predominant glutamine-consuming tissues. Under stress and/or catabolic conditions such as cancer, sepsis/infections, post-surgery and trauma, the endogenous synthesis of glutamine does not appear to meet the human body demand, and hence glutamine assumes the role of a conditionally essential amino acid(Reference Liu, Ma and Luo6Reference Altman, Stine and Dang8). The blood glutamine concentration can fall by up to 30–50 % under catabolic conditions, while at the same time, skeletal muscle glutamine concentration may fall by up to 50 %(Reference Smedberg, Rooyackers and Norberg5,Reference Cruzat, Pantaleao and Donato7,Reference Cruzat, Bittencourt and Scomazzon9) .

Glutamine has a γ-amide nitrogen that is essential for the biosynthesis of nucleotides and other metabolites such as hexosamine. In nucleotide biosynthesis, glutamine and glutamate either directly or indirectly, via aspartate for example, serve as the nitrogen donors for all nitrogen atoms in purines and pyrimidines(Reference Chowdhry, Zanca and Rajkumar10). For rapidly dividing cells such as tumour cells, enterocytes and lymphocytes, glutamine consumption may be related to an urgent need for nucleotide biosynthesis(Reference Curi, Lagranha and Doi11). Glutamine is also used for the synthesis of GSH, the major endogenous antioxidant molecule in the cell. Cells are exposed to oxidative stress under conditions such as nutrient starvation (or oversupply) and catabolic stress such as surgery, sepsis or infection, but also during cell proliferation(Reference Cruzat, Pantaleao and Donato7,Reference Cruzat, Keane and Scheinpflug12) . The antioxidant properties mediated by glutamine are discussed later in this review.

Glutamine is an important nitrogen donor for the production of NAD+, in the last steps of both the de novo (from dietary tryptophan) and Preiss–Handler (from dietary niacin) pathways(Reference Matsuyama, Yoshinaga and Shibue13). NAD+ is an essential coenzyme and electron acceptor in catabolic pathways of metabolism(Reference Ryu, Nandu and Kim14). NAD+ is also produced through salvage pathways from nicotinamide and nicotinamide riboside precursors, however, people with deficiency or low activity of the GS gene exhibit severe secondary NAD+ deficiency(Reference Hu, Ibrahim and Stucki15), indicating that the glutamine supply is indispensable for NAD+ synthesis(Reference Wojcik, Seidle and Bieganowski16).

It is widely accepted that immune cell intermediary metabolism is critical to cell function. Glucose is mainly converted into lactate (glycolysis), whereas glutamine is converted into glutamate, aspartate and alanine by undergoing partial oxidation in the TCA cycle to carbon dioxide, aspartate, alanine and lactate in a process called glutaminolysis(Reference Curi, Newsholme and Marzuca-Nassr3). Furthermore, through the pentose phosphate pathway, cells can produce ribose-5-phosphate (a five-carbon sugar), which is a precursor for the pentose sugars required for synthesis of RNA and DNA, as well as glycerol-3-phosphate for phospholipid synthesis. The degradation of glutamine, via GLS action, yields ammonia (from the amide nitrogen of glutamine) for the ATP-dependent formation of carbamoylphosphate and also the formation of aspartate through glutaminolysis which ultimately leads to the synthesis of purines and pyrimidines required for DNA and RNA synthesis(Reference Wojcik, Seidle and Bieganowski16,Reference Mills, Kelly and O'Neill17) .

Glutamine and monocyte/macrophage cell function

Macrophages are associated with high consumption rates of glutamine and in addition, glutamine is essential for many of the functions of these leucocytes. The supplementation or in vitro treatment with glutamine optimises the function of these cells(Reference Cruzat, Macedo Rogero and Noel Keane4). Macrophages are diffusely scattered in the connective tissue and in liver (where they are termed Kupffer cells), spleen and lymph nodes (sinus histiocytes), lungs (alveolar macrophages) and central nervous system (CNS) (microglia). The half-life of the macrophage precursor, the blood monocytes, is about 1 day, whereas the life span of tissue macrophages is several months or years. The mononuclear phagocyte system, consisting of a number of cell types including monocytes and macrophages, is part of both humoral and cell-mediated immunity and has an important role in defence against microorganisms, including mycobacteria, fungi, bacteria, protozoa and viruses.

With respect to the mononuclear phagocytes, different populations of macrophages have now been identified, in vivo, including M1 and M2 macrophages(Reference Martinez, Sica and Mantovani18). M1-like macrophages are responsible for secreting pro-inflammatory cytokines and lipid mediators and are involved in tissue degradation and T lymphocyte activation as part of an inflammatory immune response(Reference Gordon and Taylor19). M2 macrophages exert different functions, such as contribution to tissue repair and the secretion of anti-inflammatory cytokines and alternative lipid mediators. Lipopolysaccharides (LPS) exposure in vitro promotes a switch from glucose-dependent oxidative phosphorylation to aerobic glycolysis in macrophages – typical of the Warburg effect(Reference Gordon and Martinez20). Pyruvate kinase M2 regulates hypoxia-inducible factor 1-α activity, a rapid increase in ATP formation and IL-1β expression, thus being a key regulator of the Warburg effect in LPS-activated macrophages(Reference Palsson-McDermott, Curtis and Goel21,Reference O'Neill and Pearce22) . M1 macrophages (treated with LPS) have two points of substrate flux deviation with regards to the TCA cycle (in contrast to M2 macrophages) one occurring at the isocitrate dehydrogenase step and another post-succinate formation. This is an important metabolic reprogramming mechanism. As a result, there is an accumulation of TCA cycle intermediates (e.g. succinate, α-ketoglutarate, citrate and itaconate) that impacts the activation of LPS-stimulated macrophages(Reference Harber, de Goede and Verberk23,Reference Tannahill, Curtis and Adamik24) . Itaconate, a recently discovered metabolic regulator in macrophages has anti-inflammatory properties through activation of nuclear factor erythroid 2-related factor 2 Kelch-like (ECH)-associated protein 1 is a protein that in humans is encoded by the Keap1 gene. Keap1 has been shown to interact with Nrf2, a key regulator of antioxidant responses, required for the amelioration of oxidative stress alkylation(Reference Mills, Kelly and O'Neill17). In contrast, Liu et al.(Reference Liu, Wang and Li25) reported α-ketoglutarate, generated through glutaminolysis (glutamine → glutamate → α-ketoglutarate), promotes M2 macrophage differentiation. Macrophage metabolism varies with specific-tissue microenvironment, and this metabolic reprogramming is required for macrophage differentiation and function.

Macrophages can produce relatively large amounts of nitric oxide, especially the M1 inflammatory macrophage, via increased concentrations of the enzyme inducible nitric oxide synthase, which may be stimulated by LPS or other immune activator exposure. The activated macrophage can actually release the enzyme arginase to deplete local concentrations of arginine (indeed arginine is required for tumour growth). It has been reported that mouse peritoneal resident and Bacillus Calmette–Guerin-activated macrophages and human monocytes are capable of utilising glutamine at high rates, contain sufficient activity of the enzymes required to convert glutamine to citrulline (and subsequently citrulline to arginine via enzymes of a partial urea cycle) to account for observed rates of nitrite synthesis (which in turn is derived from derived from NO) in the absence of extracellular l-arginine, and will release nitrite when exposed to intermediates of the proposed glutamine → arginine pathway(Reference Murphy and Newsholme26). Thus, the M1 macrophage can release arginase to deplete local concentrations of arginine but ensures it has sufficient intracellular arginine via biosynthesis from glutamine.

Glutamine, the glutathione axis and heat shock protein responses

An important aspect to be considered under stress and/or catabolic conditions is the redox state of the cell, as it plays an essential role in cell fate. This can be obtained from the ratio between the intracellular concentration of glutathione disulphide (GSSG) and GSH(Reference Cruzat, Keane and Scheinpflug12). The ratio is normally expressed as [GSSG]/[GSH], and in an oxidative environment there is a resulting reduction of GSH and an increase in GSSG(Reference Desideri, Ciccarone and Ciriolo27). The inhibition of GSH recycling from GSSG does not appear to impact GSH content and cell survival. Conversely, the inhibition of the de novo synthesis of GSH dampens intracellular antioxidant defence mediated by this system and increases oxidative stress(Reference Lian, Gnanaprakasam and Wang28). The initial step for the de novo synthesis of GSH occurs through the activity of glutamate-cysteine ligase, a heterodimer of a catalytic subunit and a modifier subunit that catalyses the first and rate-limiting step to form the dipeptide γ-glutamylcysteine from cysteine and glutamate(Reference Desideri, Ciccarone and Ciriolo27).

Glutamine via glutaminolysis can increase the substrate availability for the de novo synthesis of GSH(Reference Cruzat, Bittencourt and Scomazzon9). Numerous studies in animal models submitted to infection and trauma(Reference Cruzat, Pantaleao and Donato7,Reference Cruzat, Bittencourt and Scomazzon9) , as well as critically ill and post-surgery(Reference Wang, Wu and Zheng29) patients, have demonstrated that exogenous administration of glutamine (both enteral and parenteral nutrition), significantly enhances the GSH system, increasing cellular resistance to lesions and reducing oxidative stress(Reference Cruzat, Macedo Rogero and Noel Keane4,Reference Smedberg and Wernerman30) . Interestingly, similar effects have been reported in animal models submitted to exercise. For example, chronic oral administration with glutamine effectively increased glutamine availability in tissues, such as the liver and muscle, which amplified GSH stores and improved the redox state of the cell. These and other immune-inflammatory effects, such as lowering plasma cytokines, were observed in animal models submitted to swimming(Reference Cruzat, Rogero and Tirapegui31,Reference Cruzat and Tirapegui32) , treadmill(Reference Petry, Cruzat and Heck33,Reference Petry, Cruzat and Heck34) and resistance exercise(Reference Raizel, Leite and Hypolito35,Reference Leite, Raizel and Hypolito36) .

Whether glutamine availability is maintained by its exogenous supply or endogenously synthesised, this amino acid has unique metabolic and anti-inflammatory roles. One of the main anti-inflammatory properties mediated by glutamine is through the modulation of proteins with chaperone function, such as the HSPs(Reference Leite, Cruzat and Krause37). HSPs are classified into six families (i.e. HSP10, HSP40, HSP60, HSP70, HSP90 and HSP100) on the basis of their monomeric molecular weight. Almost all HSPs prevent the aggregation of newly synthesised polypeptide chains during folding, and some HSPs, such as HSP70 have also the ability to clear improperly folded and unfolded proteins. Under several stress conditions, HSPs are primarily regulated by transcriptional activators, such as the heat shock factors (HSFs)(Reference Martinez, Dias and Natov38). HSFs undergo trimerisation and translocate to the nucleus, where they bind heat shock elements within the promoter regions of their target genes, and activate the transcription of HSPs(Reference Dimauro, Mercatelli and Caporossi39). In addition to their role as molecular chaperons, HSPs are also able to associate with the complex formed by NF-κB with its inhibitor, and block NF-κB translocation to the nucleus. Hence, the HSP response has anti-inflammatory properties by virtue of turning NF-κB off and attenuating the production of inflammatory mediators. Glutamine enhances the transactivation of HSF1 and induces HSF1 expression via cytosine-cytosine-adenosine-adenosine-thymidine (CCAAT) box motif which is present in several gene promoters enhancer-binding protein-β in a dependent manner. Furthermore, glutamine induces HSP expression via N-acetyl-O-glycosylation and phosphorylation of HSF1 and Sp1 transcription factors. This pathway is coordinated by a special and nutrient-sensing pathway, the hexosamine biosynthetic pathway(Reference Sun, Shang and Yao40,Reference Tan, Sim and Long41) . Many in vitro (Reference Cruzat, Keane and Scheinpflug12), and in vivo experiments(Reference Moura, Lollo and Morato42), including animal models submitted to exhaustive exercise training(Reference Petry, Cruzat and Heck33,Reference Petry, Cruzat and Heck34) have identified the important role of glutamine availability/metabolism and the anti-inflammatory response mediated by HSPs. Interestingly, other glutamine coupling amino acids, such as arginine have shown to stimulate HSP effects in exercised rats(Reference Moura, Lollo and Morato42).

Glutamine, the brain and exercise

Glutamine is abundantly expressed throughout the mammalian CNS, with concentrations ranging from 5 nmol/mg in the hippocampus to 7 nmol/mg in the cerebellum(Reference Takado, Sato and Kanbe43). Glutamine within the CNS functions to regulate neurotransmission through the metabolism of the neurotransmitters, glutamate and GABA(Reference Schousboe44).

Glutamate and GABA are the major excitatory and inhibitory neurotransmitters expressed within the brain and function in concert to coordinate a plethora of physiological functions from memory formation to motivation and from respiration to appetite control(Reference Xu, Bartolome and Kong45). Within the mammalian brain, evolution has shaped several brain-glutamine intrinsic ‘innovations’ which are fundamental to the efficacy and predominance of glutamate and GABA neurotransmission. The blood–brain barrier tightly regulates brain extracellular composition of glutamine, glutamate and GABA irrespective of the level in the blood(Reference Hawkins and Vina46,Reference Dolgodilina, Imobersteg and Laczko47) . As the blood–brain barrier is relativity impermeable, the brain is capable of autonomous productions of key neurotransmitters(Reference Hawkins and Vina46). Glutamatergic/GABAergic synapses are encompassed by astrocytes with excitatory amino-acid transporters and high affinity GABA transporter 1 and 3 (Fig. 2)(Reference Leenaars, Drinkenburg and Nolten48). These transporters are capable of rapidly removing glutamate and GABA from synaptic clefts post-depolarisation in order to dramatically increasing the signal-to-noise ratio of neurotransmission(Reference Boddum, Jensen and Magloire49,Reference Leke and Schousboe50) . Glutamate and GABA neurotransmitter homoeostasis is tightly regulated via glutamine through the glutamine-GGC (Fig. 2)(Reference Leenaars, Drinkenburg and Nolten48). Here, glutamine is compartmentalised not within neurons but in neighbouring astrocytes. Astrocytes exclusively express GS hence inherently limiting glutamine production to astrocytes(Reference Castegna and Menga51). This unique feature of astrocytic glutamine pooling is critical to the spatial and temporal resolution of glutamate and GABA neuronal transmission(Reference Albrecht and Zielinska52).

Fig. 2. Simplified schematic diagram of the glutamine-glutamate/GABA cycle (GGC). Within a glutamatergic and GABAergic neuron, glutamate (excitatory neurotransmitter) and GABA (inhibitory neurotransmitter) are released into the synaptic cleft. Following transmission, glutamate and GABA are taken up by excitatory amino-acid transporters (EAAT) and GABA transporter (GAT) 1 and 3 located in membranes of astrocytes. Glutamate taken up by the astrocytes is quickly aminated by glutamine synthetase (GS) to glutamine, whereas GABA is converted to glutamine via the tricarboxylic acid cycle and then to glutamine by GS. Glutamine exits astrocytes through the system N transporter 1 (SN1) transporter proteins to the extracellular space and diffuses to the surface of glutamatergic neurons, where it is taken up by diamine acetyltransferase (SAT1). Within the pre-synaptic neuron, glutamine is converted by phosphate activated glutaminase (PAG) back to glutamate or further converted to GABA by glutamic acid decarboxylase (GAD).

To replenish glutamine back to neurons, astrocytes shuttle glutamine through system N 1 transporter to diffuse across the extracellular matrix to the surface of presynaptic neurons. Here, glutamine it is taken up by diamine acetyltransferase (SAT1) and metabolised back to glutamate and GABA by the action of GLS and glutamic acid decarboxylase, respectively(Reference Verkhratsky, Nedergaard and Hertz53). Through the regulation of glutamate and GABA, glutamine within the CNS plays an integral role in many aspects of neurobiology and as such disturbances in glutamine metabolism and/or transport have been implicated in a range of pathologies ranging from chronic stress, epilepsy, hepatic encephalopathy, manganese encephalopathy, neurodegenerative disorders and metabolic disease(Reference Huang, Liu and Yin54,Reference Ivens, Caliskan and Papageorgiou55) .

A particularly unique feature of the GGC within the brain is that it is metabolically expensive. The clearance of glutamate by neurons and astrocytes from the synapse involves the co- and counter-transport of Na+ and K+ via metabolically demanding Na+–K+ pumps during neurotransmission(Reference Herbst and Holloway56). Further metabolic expenses is required for the subsequent formation of glutamine via GS in astrocytes as this is an ATP-dependent process(Reference Herbst and Holloway56). 13C-magnetic resonance spectroscopy kinetic studies tracing found that 60–80 % of resting energy utilised by both the human or rat brain is used to recycling glutamate and GABA between synapses and presynaptic terminals through the GGC(Reference Rothman, Behar and Hyder57). At rest the adult brain consumes about 20 % of the body's RMR (44–87 % during infancy, childhood and adolescence) despite accounting for only 2 % of body mass(Reference Kuzawa and Blair58). As such, GGC could account for as much as about 16 % of adult RMR indicating that glutamine metabolism is vulnerable during periods of metabolic activity such as exercise.

An emerging body of evidence suggests that physical exercise has a profound influence on the CNS. Pre-clinical and human studies demonstrate that exercise, whether that be acute or chronic, regulates neuronal activity particularly in regions such as the hippocampus, hypothalamus, cerebellum, basal ganglia and motor cortex to influence the neurophysiology underlying memory, neurogenesis, motivation and appetite(Reference Seo59Reference Rodrigues, Pereira and de Campos63).

Unlike whole-body energy balance (balance between energy intake and expenditure), global brain energy balance is not altered during low-intensity exercise or during the transition from rest to exercise, however, increased metabolic rate is detectable at the whole-brain level during vigorous exercise(Reference Smith and Ainslie64,Reference Coxon, Cash and Hendrikse65) . Maddock et al. reported that vigorous exercise in human subjects increases glutamine levels within the cortex, an effect that occurs as a consequence of activity-dependent increases in the glutamate–glutamine turnover(Reference Maddock, Casazza and Buonocore66). Furthermore, involuntary exercise to exhaustion in rats, increased glutamine in the hippocampus, striatum and cerebellum(Reference Swiatkiewicz, Fiedorowicz and Orzel67). As exercise has been shown to attenuate the elevated central glutamate transmission in models of addiction and cardiovascular functions, it remains unclear if elevated levels of glutamine are a cause or consequence to these changes in brain function. It is well established that exhaustive exercise increases circulating NH3(Reference Weiner and Verlander68), and one possible explanation for the enhanced glutamine within the brain could be to detoxify the brain via the removal of NH3 through the synthesis of glutamine using GS(Reference Conway and Hutson69). Interestingly, branched chain amino acids (i.e. leucine, isoleucine, valine), glutamine and arginine may hypothetically enhance exercise performance and decrease serotonin, a mediator involved in central fatigue(Reference Conway and Hutson69). However, there is lack of in vivo evidence demonstrating that amino acid supplementation may effectively reduce central fatigue. The relative contribution of glutamine to the enhanced performance reported in relatively few studies has not yet been confirmed thus care must be taken with any glutamine supplementation exercise particularly as elevated GS levels are described to be elevated in rodent models of diet-induced obesity(Reference Soontornniyomkij, Kesby and Soontornniyomkij70) and other pathological conditions(Reference Huyghe, Denninger and Voss71). In addition, it is well known that high levels of NH3 induced by high amino acid intake have toxic effects, especially for the brain and the kidneys(Reference Conway and Hutson69,Reference Davani-Davari, Karimzadeh and Sagheb72) . Although there is an emerging role of central glutamine metabolism in exercise, the relative contributions of the GGC and its molecular machinery (GS, system N 1, SAT1, GLS and glutamic acid decarboxylase) to enhance exercise performance, including exercise-induced regulation of neuronal function and/or central fatigue remains unclear.

Glutamine supplementation

The relationship between glutamine metabolism (plasma and tissue), supplementation and immune-inflammatory responses have been investigated for more than 35 years(Reference Curi, Newsholme and Marzuca-Nassr3,Reference Cruzat, Pantaleao and Donato7) . The literature demonstrates an overall consensus regarding the fall of glutamine in critically ill patients, and this has been driving studies that provide glutamine exogenous supply. However, in situations where catabolism is more dynamic and/or less severe, which include exercise situations, glutamine levels in the body may fluctuate, and hence results from research might be conflicting and/or lacking in the literature (Fig. 1). For instance, Borgenvik et al.(Reference Borgenvik, Nordin and Mikael Mattsson73) observed a reduction in plasma concentrations of essential amino acids and glutamine during an adventure racing exercise, whereas no changes were detected in muscle concentration after exercise. Similar findings were also reported by Tritto et al.(Reference Tritto, Amano and De Cillo74) where plasma glutamine was unchanged and not associated with immunosuppressive responses in combat athletes undergoing to a rapid weight loss.

It is worth emphasising that, although plasma glutamine level is reduced from its normal concentration (i.e. 500–800 μmol/l) to 300–400 μmol/l, cells depending on this amino acid, are negatively impacted in terms of proliferation and function by this decrease(Reference Cruzat, Krause and Newsholme75). Exhaustive exercise can dramatically increase the catabolism of amino acids in the skeletal muscles and liver, which eventually might reduce the concentration of glutamine in plasma and associated tissues. However, this is more likely to be a gluconeogenic transitory situation, as glutamine provides nitrogen atoms to the synthesis of purines, pyrimidines and amino sugars. If the high glutamine degradation persists, many metabolic pathways and mechanisms that depend on glutamine availability can be affected and increase the potential of immunosuppression. Indeed, several studies have reported that immunosuppression induced by exercise excess can increase the risk of infections and the development or aggravation of overtraining syndrome(Reference Gleeson, Pyne and Elkington76,Reference Edouard, Junge and Sorg77) . However, the link between glutamine availability and infection seems to be clearer in critically ill patients(Reference Smedberg and Wernerman30) and less likely to play a mechanistic role in exercise-induced immunodepression(Reference Ramezani Ahmadi, Rayyani and Bahreini78).

Scientific evidence has shown the important immune-inflammatory roles of glutamine availability and/or glutamine metabolism in stressful and/or catabolic situations. However, several questions and specific considerations were raised in regards to quantity/dose, frequency and form of administration of l-glutamine supplementation. l-Glutamine is often supplied as a component of clinical nutrition supplementation pre- and post-operative surgery or in critically ill patients. In these situations, l-glutamine is usually administered by utilising its free form (also known as an isolated amino acid), or bond with another amino acid, also known as the dipeptide form, such as l-alanyl-l-glutamine or l-glycine-l-glutamine. In general, many clinical and experimental studies have shown more beneficial results when glutamine was administered as an intravenous solution, when compared to enteral routes. For instance, intravenous l-glutamine administration appears to reduce the rate of infectious complications, length of hospital stay and mortality of critically ill patients(Reference Smedberg and Wernerman30,Reference Stehle, Ellger and Kojic79,Reference McRae80) . It is important to note that these results are more likely to be achieved and/or be less heterogenous among patients and studies where intravenous or parenteral l-glutamine is provided to individuals with or without hypoglutaminaemia and receiving adequate energetic and protein supply (e.g. according to ESPEN recommendations(Reference Cruzat, Macedo Rogero and Noel Keane4,Reference Smedberg and Wernerman30) ). Conversely, intravenous or parenteral routes are always invasive, and it cannot be applied for sport and exercise situations. If l-glutamine supplementation is eventually recommended for elite athletes, oral routes are the best choice, although it is well known that rapidly dividing cells, such as the enterocytes, are extremely avid consumers of the amino acid. Glutamine is absorbed in the upper part of the small intestine and subsequently metabolised in the liver, thus it may not be available in sufficient quantity for other tissue targets, such as immune and muscle cells, to have a measurable effect.

An alternative way to bypass the high glutamine consumption of enterocytes is oral supplementation with glutamine dipeptides. Bonded amino acids (e.g. di- and tripeptides) have a differentiated membrane transport in the luminal membrane through the glycopeptide transport protein-1(Reference Boudry, Rome and Perrier81). This system allows a higher proportion of di- and tripeptides to escape hydrolysis and metabolisation be the enterocytes, and hence are more efficiently transported into the bloodstream and other tissues, such as the liver, immune system, kidneys and skeletal muscles. Interestingly, when free l-glutamine and free l-alanine are administered orally in combination, similar immune inflammatory effects were observed in animal models submitted to exhaustive exercise(Reference Cruzat and Tirapegui32). Although the precise mechanisms are still unknown, l-glutamine and l-alanine work in parallel, i.e. l-alanine is rapidly metabolised via alanine aminotransferase to pyruvate, with concomitant production of glutamate from 2-oxoglutarate, which contribute to antioxidant defence, such as GSH production. Hence, the presence of l-alanine in the l-alanyl-l-glutamine dipeptide or in its free form and supplemented in tandem, can spare glutamine metabolism in order for the latter to be used by high-demand tissues under inflammatory and/or highly catabolic situations(Reference Cruzat, Pantaleao and Donato7,Reference Cruzat and Tirapegui32) . This promotes the idea that l-glutamine supplementation should be accompanied by another amino acids or possibly proteins. Indeed, when l-glutamine was administered with proteins, such as wheat protein, or branched chain amino acids the whole-body glutamine status was increased(Reference Harris, Hoffman and Allsopp82) and muscle atrophy was more effectively inhibited(Reference Haba, Fujimura and Oyama83). Moreover, other amino acids, such as arginine may provide additional effects upon defence mechanisms, including the HSP response(Reference Moura, Lollo and Morato42). If the efficacy of l-glutamine supplementation is dramatically impacted by the presence or absence of other amino acids metabolised by the enterocytes, this would possibly explain, at least in part controversial results obtained in human subjects and animal models supplemented with l-glutamine and submitted to exercise. Other reasons for heterogeneous results include the quantity/dose and frequency of l-glutamine supplementation. At least in experimental models with rodents, both factors can significantly impact the possible immune and antioxidant effects mediated by l-glutamine supplementation.

Overall, current data indicate that glutamine supplementation may provide important metabolic and immunological support in catabolic situations, including critical care and elite athletes. However, as reported by other authors(Reference Smedberg and Wernerman30), several precise scientific questions, hypothesis and mechanistic explanations about the effects of glutamine exogenous supply were never tested in critically ill patients, nor in sport and exercise. For instance, should hypoglutaminaemia and/or tissue glutamine stores be the main factor driving the decision of supply of exogenous glutamine? Also, in catabolic situations that hypoglutaminaemia is not observed, can individuals obtain benefits, such as immune and antioxidant effects from glutamine supplementation?

Conclusion

The scientific literature continues to provide evidence of the importance of glutamine metabolism in cell function and homoeostasis, especially via heat shock responses and the GSH system. This is particularly important for cells of the immune system, such as monocytes/macrophages to maintain a continuous defence against microorganisms. Conversely, several questions remain about the efficacy of l-glutamine supplementation in catabolic situations, such as critical care and extenuating exercise. Possibly, the lack of consensus is related to the fact that quantity/dose and frequency of l-glutamine supplementation, as well as the presence or absence of other amino acids metabolised by the enterocytes significantly impact the desired amino acid supplementation outcomes. In the CNS for example, while disturbances in glutamine metabolism have been implicated in a range of neurodegenerative disorders, as glutamine is vital for glutamate and GABA concentrations, high levels of NH3 induced by high amino acid intake may have toxic effects. Although in vitro and animal studies continue to be important for the understanding of cell metabolism and physiology, more studies in human subjects are required to better elucidate the action and positive benefits of glutamine supplementation. The following suggestions may aid in deciding the formulation of future studies in immunometabolism:

  • Determination of the role of glutamine metabolism and exogenous supply of glutamine-derived metabolic intermediates in the essential antiviral activity of immune cells.

  • Glutamine supplementation and potential slowing of immunosenescence by altering immune cell metabolism and function.

  • Mechanistic studies on enteral absorption of glutamine supplements, including concomitant amino acid administration to optimise glutamine immune-metabolic effects.

Acknowledgements

We acknowledge the Faculty of Health, Torrens University Australia for research support. We also acknowledge the Curtin Medical School and The Curtin Health Innovation Research Institute for research support and provision of excellent research facilities. G. T. D. also acknowledge The University of Melbourne for provision of research facilities.

Financial Support

This study was supported by Australian Research Council (ARC) Linkage grant LP190100130, National Health and Medical Research Council (NHMRC) of Australia (grants: 1160043, 2002427), Diabetes Australia Trust (grant: Y20G-DODG) and the Australian Research Council for research funding (grant: DP220102132) and the São Paulo Research Foundation (FAPESP), grant no. 2021/13119-2.

Conflict of Interest

The authors declare no conflicting or competing interests.

Authorship

P. N. and V. C. were responsible for the review outline and decisions regarding content of the review. V. C. prepared figures associated with this review. All authors contributed to writing of the text.

References

Wu, D, Lewis, ED, Pae, M et al. (2018) Nutritional modulation of immune function: analysis of evidence, mechanisms, and clinical relevance. Front Immunol 9, 3160.10.3389/fimmu.2018.03160CrossRefGoogle ScholarPubMed
Calder, PC, Carr, AC, Gombart, AF et al. (2020) Optimal nutritional status for a well-functioning immune system is an important factor to protect against viral infections. Nutrients 12, 1181.CrossRefGoogle ScholarPubMed
Curi, R, Newsholme, P, Marzuca-Nassr, GN et al. (2016) Regulatory principles in metabolism-then and now. Biochem J 473, 18451857.CrossRefGoogle ScholarPubMed
Cruzat, V, Macedo Rogero, M, Noel Keane, K et al. (2018) Glutamine: metabolism and immune function, supplementation and clinical translation. Nutrients 10, 1564.CrossRefGoogle ScholarPubMed
Smedberg, M, Rooyackers, O, Norberg, Å et al. (2020) Endogenous production of glutamine and plasma glutamine concentration in critically ill patients. Clin Nutr ESPEN 40, 226230.CrossRefGoogle ScholarPubMed
Liu, N, Ma, X, Luo, X et al. (2018) l-Glutamine attenuates apoptosis in porcine enterocytes by regulating glutathione-related redox homeostasis. J Nutr 148, 526534.CrossRefGoogle ScholarPubMed
Cruzat, VF, Pantaleao, LC, Donato, J Jr et al. (2014) Oral supplementations with free and dipeptide forms of l-glutamine in endotoxemic mice: effects on muscle glutamine-glutathione axis and heat shock proteins. J Nutr Biochem 25, 345352.CrossRefGoogle ScholarPubMed
Altman, BJ, Stine, ZE & Dang, CV (2016) From Krebs to clinic: glutamine metabolism to cancer therapy. Nat Rev Cancer 16, 619634.CrossRefGoogle ScholarPubMed
Cruzat, VF, Bittencourt, A, Scomazzon, SP et al. (2014) Oral free and dipeptide forms of glutamine supplementation attenuate oxidative stress and inflammation induced by endotoxemia. Nutrition 30, 602611.CrossRefGoogle ScholarPubMed
Chowdhry, S, Zanca, C, Rajkumar, U et al. (2019) NAD metabolic dependency in cancer is shaped by gene amplification and enhancer remodelling. Nature 569, 570575.CrossRefGoogle ScholarPubMed
Curi, R, Lagranha, CJ, Doi, SQ et al. (2005) Molecular mechanisms of glutamine action. J Cell Physiol 204, 392401.CrossRefGoogle ScholarPubMed
Cruzat, VF, Keane, KN, Scheinpflug, AL et al. (2015) Alanyl-glutamine improves pancreatic beta-cell function following ex vivo inflammatory challenge. J Endocrinol 224, 261271.CrossRefGoogle ScholarPubMed
Matsuyama, T, Yoshinaga, SK, Shibue, K et al. (2021) Comorbidity-associated glutamine deficiency is a predisposition to severe COVID-19. Cell Death Differ 28, 31993213.CrossRefGoogle ScholarPubMed
Ryu, KW, Nandu, T, Kim, J et al. (2018) Metabolic regulation of transcription through compartmentalized NAD(+) biosynthesis. Science (New York, NY) 360, eaan5780.CrossRefGoogle ScholarPubMed
Hu, L, Ibrahim, K, Stucki, M et al. (2015) Secondary NAD+ deficiency in the inherited defect of glutamine synthetase. J Inherit Metab Dis 38, 10751083.CrossRefGoogle ScholarPubMed
Wojcik, M, Seidle, HF, Bieganowski, P et al. (2006) Glutamine-dependent NAD+ synthetase. How a two-domain, three-substrate enzyme avoids waste. J Biol Chem 281, 3339533402.CrossRefGoogle ScholarPubMed
Mills, EL, Kelly, B & O'Neill, LAJ (2017) Mitochondria are the powerhouses of immunity. Nat Immunol 18, 488498.CrossRefGoogle ScholarPubMed
Martinez, FO, Sica, A, Mantovani, A et al. (2008) Macrophage activation and polarization. Frontiers in Bioscience: A Journal and Virtual Library 13, 453461.CrossRefGoogle ScholarPubMed
Gordon, S & Taylor, PR (2005) Monocyte and macrophage heterogeneity. Nat Rev Immunol 5, 953964.CrossRefGoogle ScholarPubMed
Gordon, S & Martinez, FO (2010) Alternative activation of macrophages: mechanism and functions. Immunity 32, 593604.CrossRefGoogle ScholarPubMed
Palsson-McDermott, EM, Curtis, AM, Goel, G et al. (2015) Pyruvate kinase M2 regulates Hif-1α activity and IL-1β induction and is a critical determinant of the Warburg effect in LPS-activated macrophages. Cell Metab 21, 6580.CrossRefGoogle ScholarPubMed
O'Neill, LA & Pearce, EJ (2016) Immunometabolism governs dendritic cell and macrophage function. J Exp Med 213, 1523.CrossRefGoogle ScholarPubMed
Harber, KJ, de Goede, KE, Verberk, SGS et al. (2020) Succinate is an inflammation-induced immunoregulatory metabolite in macrophages. Metabolites 10, 372.CrossRefGoogle ScholarPubMed
Tannahill, GM, Curtis, AM, Adamik, J et al. (2013) Succinate is an inflammatory signal that induces IL-1β through HiF-1α. Nature 496, 238242.CrossRefGoogle ScholarPubMed
Liu, PS, Wang, H, Li, X et al. (2017) α-Ketoglutarate orchestrates macrophage activation through metabolic and epigenetic reprogramming. Nat Immunol 18, 985994.CrossRefGoogle ScholarPubMed
Murphy, C & Newsholme, P (1998) Importance of glutamine metabolism in murine macrophages and human monocytes to l-arginine biosynthesis and rates of nitrite or urea production. Clin Sci 95, 397407.CrossRefGoogle ScholarPubMed
Desideri, E, Ciccarone, F & Ciriolo, MR (2019) Targeting glutathione metabolism: partner in crime in anticancer therapy. Nutrients 11, 1926.CrossRefGoogle ScholarPubMed
Lian, G, Gnanaprakasam, JR, Wang, T et al. (2018) Glutathione de novo synthesis but not recycling process coordinates with glutamine catabolism to control redox homeostasis and directs murine T cell differentiation. eLife 7, e36158.CrossRefGoogle Scholar
Wang, ZE, Wu, D, Zheng, LW et al. (2018) Effects of glutamine on intestinal mucus barrier after burn injury. Am J Transl Res 10, 38333846.Google ScholarPubMed
Smedberg, M & Wernerman, J (2016) Is the glutamine story over? Crit Care 20, 361.CrossRefGoogle ScholarPubMed
Cruzat, VF, Rogero, MM & Tirapegui, J (2010) Effects of supplementation with free glutamine and the dipeptide alanyl-glutamine on parameters of muscle damage and inflammation in rats submitted to prolonged exercise. Cell Biochem Funct 28, 2430.CrossRefGoogle ScholarPubMed
Cruzat, VF & Tirapegui, J (2009) Effects of oral supplementation with glutamine and alanyl-glutamine on glutamine, glutamate, and glutathione status in trained rats and subjected to long-duration exercise. Nutrition 25, 428435.CrossRefGoogle ScholarPubMed
Petry, ER, Cruzat, VF, Heck, TG et al. (2015) l-Glutamine supplementations enhance liver glutamine-glutathione axis and heat shock factor-1 expression in endurance-exercise trained rats. Int J Sport Nutr Exerc Metab 25, 188197.CrossRefGoogle ScholarPubMed
Petry, ER, Cruzat, VF, Heck, TG et al. (2014) Alanyl-glutamine and glutamine plus alanine supplements improve skeletal redox status in trained rats: involvement of heat shock protein pathways. Life Sci 94, 130136.CrossRefGoogle ScholarPubMed
Raizel, R, Leite, JS, Hypolito, TM et al. (2016) Determination of the anti-inflammatory and cytoprotective effects of l-glutamine and l-alanine, or dipeptide, supplementation in rats submitted to resistance exercise. Br J Nutr 116, 470479.CrossRefGoogle ScholarPubMed
Leite, JS, Raizel, R, Hypolito, TM et al. (2016) l-Glutamine and l-alanine supplementation increase glutamine-glutathione axis and muscle HSP-27 in rats trained using a progressive high-intensity resistance exercise. Appl Physiol Nutr Metab 41, 842849.CrossRefGoogle ScholarPubMed
Leite, JSM, Cruzat, VF, Krause, M et al. (2016) Physiological regulation of the heat shock response by glutamine: implications for chronic low-grade inflammatory diseases in age-related conditions. Nutrire 41, 17.CrossRefGoogle Scholar
Martinez, MR, Dias, TB, Natov, PS et al. (2017) Stress-induced O-GlcNAcylation: an adaptive process of injured cells. Biochem Soc Trans 45, 237249.CrossRefGoogle ScholarPubMed
Dimauro, I, Mercatelli, N & Caporossi, D (2016) Exercise-induced ROS in heat shock proteins response. Free Radical Biol Med 98, 4655.CrossRefGoogle ScholarPubMed
Sun, C, Shang, J, Yao, Y et al. (2016) O-GlcNAcylation: a bridge between glucose and cell differentiation. J Cell Mol Med 20, 769–781.CrossRefGoogle ScholarPubMed
Tan, HWS, Sim, AYL & Long, YC (2017) Glutamine metabolism regulates autophagy-dependent mTORC1 reactivation during amino acid starvation. Nat Commun 8, 338.CrossRefGoogle ScholarPubMed
Moura, CS, Lollo, PCB, Morato, PN et al. (2017) Modulatory effects of arginine, glutamine and branched-chain amino acids on heat shock proteins, immunity and antioxidant response in exercised rats. Food Funct 8, 32283238.CrossRefGoogle ScholarPubMed
Takado, Y, Sato, N, Kanbe, Y et al. (2019) Association between brain and plasma glutamine levels in healthy young subjects investigated by MRS and LC/MS. Nutrients 11, 1649.CrossRefGoogle ScholarPubMed
Schousboe, A (2019) Metabolic signaling in the brain and the role of astrocytes in control of glutamate and GABA neurotransmission. Neurosci Lett 689, 1113.CrossRefGoogle ScholarPubMed
Xu, J, Bartolome, CL & Kong, D (2018) Synaptic regulation of metabolism. Adv Exp Med Biol 1090, 4977.CrossRefGoogle ScholarPubMed
Hawkins, RA & Vina, JR (2016) How glutamate is managed by the blood–brain barrier. Biology (Basel) 5, 37.Google ScholarPubMed
Dolgodilina, E, Imobersteg, S, Laczko, E et al. (2016) Brain interstitial fluid glutamine homeostasis is controlled by blood–brain barrier SLC7A5/LAT1 amino acid transporter. J Cereb Blood Flow Metab 36, 19291941.CrossRefGoogle ScholarPubMed
Leenaars, CHC, Drinkenburg, WHP, Nolten, C et al. (2019) Sleep and microdialysis: an experiment and a systematic review of histamine and several amino acids. J Circadian Rhythms 17, 7.CrossRefGoogle Scholar
Boddum, K, Jensen, TP, Magloire, V et al. (2016) Astrocytic GABA transporter activity modulates excitatory neurotransmission. Nat Commun 7, 13572.CrossRefGoogle ScholarPubMed
Leke, R & Schousboe, A (2016) The glutamine transporters and their role in the glutamate/GABA-glutamine cycle. Adv Neurobiol 13, 223257.CrossRefGoogle ScholarPubMed
Castegna, A & Menga, A (2018) Glutamine synthetase: localization dictates outcome. Genes (Basel) 9, 108.CrossRefGoogle ScholarPubMed
Albrecht, J & Zielinska, M (2019) Exchange-mode glutamine transport across CNS cell membranes. Neuropharmacology 15, 107560.CrossRefGoogle ScholarPubMed
Verkhratsky, A, Nedergaard, M & Hertz, L (2015) Why are astrocytes important? Neurochem Res 40, 389401.CrossRefGoogle ScholarPubMed
Huang, D, Liu, D, Yin, J et al. (2017) Glutamate-glutamine and GABA in brain of normal aged and patients with cognitive impairment. Eur Radiol 27, 26982705.CrossRefGoogle ScholarPubMed
Ivens, S, Caliskan, G, Papageorgiou, I et al. (2019) Persistent increase in ventral hippocampal long-term potentiation by juvenile stress: a role for astrocytic glutamine synthetase. Glia 67, 22792293.CrossRefGoogle Scholar
Herbst, EA & Holloway, GP (2016) Exercise increases mitochondrial glutamate oxidation in the mouse cerebral cortex. Appl Physiol Nutr Metab 41, 799801.CrossRefGoogle ScholarPubMed
Rothman, DL, Behar, KL, Hyder, F et al. (2003) In vivo NMR studies of the glutamate neurotransmitter flux and neuroenergetics: implications for brain function. Annu Rev Physiol 65, 401427.CrossRefGoogle ScholarPubMed
Kuzawa, CW & Blair, C (2019) A hypothesis linking the energy demand of the brain to obesity risk. Proc Natl Acad Sci USA 116, 1326613275.CrossRefGoogle ScholarPubMed
Seo, JH (2018) Treadmill exercise alleviates stress-induced anxiety-like behaviors in rats. J Exerc Rehabil 14, 724730.CrossRefGoogle ScholarPubMed
Alkadhi, KA (2018) Exercise as a positive modulator of brain function. Mol Neurobiol 55, 31123130.CrossRefGoogle ScholarPubMed
Do, K, Laing, BT, Landry, T et al. (2018) The effects of exercise on hypothalamic neurodegeneration of Alzheimer's disease mouse model. PLoS One 13, e0190205.CrossRefGoogle ScholarPubMed
He, Z, Gao, Y, Alhadeff, AL et al. (2018) Cellular and synaptic reorganization of arcuate NPY/AgRP and POMC neurons after exercise. Mol Metab 18, 107119.CrossRefGoogle ScholarPubMed
Rodrigues, K, Pereira, RM, de Campos, TDP et al. (2018) The role of physical exercise to improve the browning of white adipose tissue via POMC neurons. Front Cell Neurosci 12, 88.CrossRefGoogle ScholarPubMed
Smith, KJ & Ainslie, PN (2017) Regulation of cerebral blood flow and metabolism during exercise. Exp Physiol 102, 13561371.CrossRefGoogle ScholarPubMed
Coxon, JP, Cash, RFH, Hendrikse, JJ et al. (2018) GABA concentration in sensorimotor cortex following high-intensity exercise and relationship to lactate levels. J Physiol 596, 691702.CrossRefGoogle ScholarPubMed
Maddock, RJ, Casazza, GA, Buonocore, MH et al. (2011) Vigorous exercise increases brain lactate and Glx (glutamate + glutamine): a dynamic 1H-MRS study. NeuroImage 57, 13241330.CrossRefGoogle ScholarPubMed
Swiatkiewicz, M, Fiedorowicz, M, Orzel, J et al. (2017) Increases in brain (1)H-MR glutamine and glutamate signals following acute exhaustive endurance exercise in the rat. Front Physiol 8, 19.CrossRefGoogle ScholarPubMed
Weiner, ID & Verlander, JW (2019) Emerging features of ammonia metabolism and transport in acid–base balance. Semin Nephrol 39, 394405.CrossRefGoogle ScholarPubMed
Conway, ME & Hutson, SM (2016) BCAA metabolism and NH3 homeostasis. Adv Neurobiol 13, 99132.CrossRefGoogle ScholarPubMed
Soontornniyomkij, V, Kesby, JP, Soontornniyomkij, B et al. (2016) Age and high-fat diet effects on glutamine synthetase immunoreactivity in liver and hippocampus and recognition memory in mice. Curr Aging Sci 9, 301309.CrossRefGoogle ScholarPubMed
Huyghe, D, Denninger, AR, Voss, CM et al. (2019) Phosphorylation of glutamine synthetase on threonine 301 contributes to its inactivation during epilepsy. Front Mol Neurosci 12, 120.CrossRefGoogle ScholarPubMed
Davani-Davari, D, Karimzadeh, I, Sagheb, MM et al. (2019) The renal safety of l-carnitine, l-arginine, and glutamine in athletes and bodybuilders. J Renal Nutr 29, 221234.CrossRefGoogle ScholarPubMed
Borgenvik, M, Nordin, M, Mikael Mattsson, C et al. (2012) Alterations in amino acid concentrations in the plasma and muscle in human subjects during 24 h of simulated adventure racing. Eur J Appl Physiol 112, 36793688.CrossRefGoogle ScholarPubMed
Tritto, ACC, Amano, MT, De Cillo, ME et al. (2018) Effect of rapid weight loss and glutamine supplementation on immunosuppression of combat athletes: a double-blind, placebo-controlled study. J Exerc Rehabil 14, 8392.CrossRefGoogle ScholarPubMed
Cruzat, VF, Krause, M & Newsholme, P (2014) Amino acid supplementation and impact on immune function in the context of exercise. J Int Soc Sports Nutr 11, 61.CrossRefGoogle ScholarPubMed
Gleeson, M, Pyne, DB, Elkington, LJ et al. (2017) Developing a multi-component immune model for evaluating the risk of respiratory illness in athletes. Exerc Immunol Rev 23, 5264.Google ScholarPubMed
Edouard, P, Junge, A, Sorg, M et al. (2019) Illnesses during 11 international athletics championships between 2009 and 2017: incidence, characteristics and sex-specific and discipline-specific differences. Br J Sports Med 53, 11741182.CrossRefGoogle ScholarPubMed
Ramezani Ahmadi, A, Rayyani, E, Bahreini, M et al. (2019) The effect of glutamine supplementation on athletic performance, body composition, and immune function: a systematic review and a meta-analysis of clinical trials. Clin Nutr 38, 10761091.CrossRefGoogle Scholar
Stehle, P, Ellger, B, Kojic, D et al. (2017) Glutamine dipeptide-supplemented parenteral nutrition improves the clinical outcomes of critically ill patients: a systematic evaluation of randomised controlled trials. Clin Nutr ESPEN 17, 7585.CrossRefGoogle ScholarPubMed
McRae, MP (2017) Therapeutic benefits of glutamine: an umbrella review of meta-analyses. Biomed Rep 6, 576584.CrossRefGoogle ScholarPubMed
Boudry, G, Rome, V, Perrier, C et al. (2014) A high-protein formula increases colonic peptide transporter 1 activity during neonatal life in low-birth-weight piglets and disturbs barrier function later in life. Br J Nutr 112, 10731080.CrossRefGoogle ScholarPubMed
Harris, RC, Hoffman, JR, Allsopp, A et al. (2012) l-Glutamine absorption is enhanced after ingestion of l-alanylglutamine compared with the free amino acid or wheat protein. Nutr Res 32, 272277.CrossRefGoogle ScholarPubMed
Haba, Y, Fujimura, T, Oyama, K et al. (2019) Effect of oral branched-chain amino acids and glutamine supplementation on skeletal muscle atrophy after total gastrectomy in rat model. J Surg Res 243, 281288.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. Blood glutamine concentration changes according to the balance between major organ producers and consumers in health and catabolic situations. As the most abundant and versatile amino acid in the body, glutamine largely depends on the balance between its synthesis, release and uptake by organs and tissues. In turn, multiple intracellular pathways require glutamine as a substrate to maintain homoeostasis. In health, there is a balance between glutamine synthesis and degradation, while in catabolic situations organs responsible for glutamine synthesis reduce its production, such as the skeletal muscle tissue. Other glutamine producers, such as the adipose tissue and the lungs, do not have the capability to replenish the needs of the amino acid during catabolism. Moreover, the liver, a main glutamine producer in health becomes a major glutamine consumer under disease conditions, due to gluconeogenesis support. At the same time, cells of the immune system increase their demand for both glutamine and glucose. Although the brain and kidneys may have their glutamine capabilities altered according to the type of disease/catabolic condition, no significant changes may counteract the fall in glutamine (GLN) availability.

Figure 1

Fig. 2. Simplified schematic diagram of the glutamine-glutamate/GABA cycle (GGC). Within a glutamatergic and GABAergic neuron, glutamate (excitatory neurotransmitter) and GABA (inhibitory neurotransmitter) are released into the synaptic cleft. Following transmission, glutamate and GABA are taken up by excitatory amino-acid transporters (EAAT) and GABA transporter (GAT) 1 and 3 located in membranes of astrocytes. Glutamate taken up by the astrocytes is quickly aminated by glutamine synthetase (GS) to glutamine, whereas GABA is converted to glutamine via the tricarboxylic acid cycle and then to glutamine by GS. Glutamine exits astrocytes through the system N transporter 1 (SN1) transporter proteins to the extracellular space and diffuses to the surface of glutamatergic neurons, where it is taken up by diamine acetyltransferase (SAT1). Within the pre-synaptic neuron, glutamine is converted by phosphate activated glutaminase (PAG) back to glutamate or further converted to GABA by glutamic acid decarboxylase (GAD).