About all

Glutamate disorders: Glutamate: The Master Neurotransmitter and Its Implications in Chronic Stress and Mood Disorders

Glutamate: The Master Neurotransmitter and Its Implications in Chronic Stress and Mood Disorders

The Sudden Popularity of Glutamate

Until recently, glutamate has often been mentioned only as a sidenote to the more well-known neurotransmitters such as serotonin and norepinephrine. Like the shy kid who suddenly became visible with a new haircut, glutamate has taken the neuroscience literature by storm. This brief review article will explain why glutamate is deserving of this newfound attention and may well be the master neurotransmitter responsible for shaping the entire brain.

Functions and Mechanisms of Glutamate

Storage and Transmission

Over the past three decades, researchers have learned that glutamate is the major excitatory neurotransmitter of the healthy mammalian brain, as the most profuse free amino acid that happens to sit at the intersection between several metabolic pathways (Watkins and Jane, 2006; Zhou and Danbolt, 2014). Glutamate is stored in synaptic vesicles of nerve terminals until it is released by exocytosis into the extracellular fluid, where it can quickly become highly concentrated (Zhou and Danbolt, 2014). Additionally, micromolar concentrations of basal extracellular glutamate, originating from non-vesicular release from the cystine-glutamate antiporter, continue to circulate in the space outside the synaptic cleft (Baker et al., 2002). Maintaining optimal levels in this space is essential, as low levels can deplete energy whereas excess levels can lead to cell death (Zhou and Danbolt, 2014). Glutamate transporters located on the outside of astrocytes and neurons quickly act to remove excess glutamate (Zhou and Danbolt, 2014). Receptor proteins at the surface of cells detect glutamate in the extracellular fluid and receive it (Zhou and Danbolt, 2014).

Most cells in the central nervous system (CNS) express at least one type of glutamate receptor. These include the ionotropic N-methyl-D-aspartate (NMDA), AMPA (a-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid), and kainite receptors, which mediate fast excitatory transmission; in addition to the family of eight metabotropic glutamate receptors (mGluR1-8), which are located pre-, post-, and extra-syntactically throughout the CNS (Watkins and Jane, 2006; Reznikov et al. , 2011; Zhou and Danbolt, 2014). The complex and widespread mechanisms of transmission mean that there is almost unlimited potential for research on each class of receptors and sub-receptors (Watkins and Jane, 2006).

Neuroplasticity

As a neurostimulator, there is strong support for a role for glutamate in a variety of neuroplasticity mechanisms including long-term potentiation (LTP), regulation of spine density, and synaptic reorganization (Reznikov et al., 2011). As a result, glutamate is now known to be exceptionally important in cognition, learning and mood, all areas in which neuroplasticity is essential to adapting to environmental stressors (Reznikov et al., 2011). LTP in several structures of the CNS employs NMDA and AMPA glutamate receptors to strengthen synaptic connections, necessary for learning and memory (Lynch, 2004; Sah et al., 2008). Morphologic adaptation is necessary for the regulation of mood and cognition (Reznikov et al., 2011).

However, chronic stress can lead to malfunctioning of the glutamate system and reduced neuroplasticity. In the hippocampus, chronic stress leads to increased glutamate release, impaired LTP, atrophy of the apical dendrites, and learning and memory deficits (Reznikov et al., 2011). In the prefrontal cortex, chronic stress leads to decreased glutamate release, impaired LTP, reduced dendritic spines, and impaired attention (Reznikov et al., 2011). In the amygdala, chronic stress leads to decreased glutamate release, impaired or enhanced LTP, dendritic hypertrophy, increased dendritic spines, and anxiety (Reznikov et al., 2011). Guo et al. (2020) have suggested that the negative impact of stress may be due to activation of the microglial cells, which trigger neuroinflammation, affecting both intracellular and extracellular signaling pathways.

Potential for Future Treatment

Antidepressant Medications

Glutamate system dysfunction has been implicated in several pre-clinical and clinical studies of mood and disorders. Glutamate reductions have been noted in several neural areas of patients with MDD (Arnone et al. , 2015), while mixed results were found with bipolar disorder (Gigante et al., 2012; Chitty et al., 2013), and several glutamatergic genes affecting different kinds of receptors have been implicated in mood disorders (de Sousa et al., 2017). Several glutamatergic agents have been demonstrated to effectively decrease depressive symptoms in people with MDD and bipolar disorder (BD) (Henter et al., 2018).

Among the most studied is ketamine, which rapidly achieves its antidepressant effects with long-lasting effects of a small dose in even treatment resistant MDD and BD (Kantrowitz et al., 2015; Newport et al., 2015; Mandal et al., 2019). Although the mechanisms of ketamine’s actions are still not understood, preclinical studies in mice suggest that found that its antidepressant effects may be produced by the metabolite (2R,6R)-hydroxynorketamine (HNK) that increases AMPA receptor activation (Zanos et al., 2016). Intravenous esketamine, an S(+) enantiomer of ketamine with a high affinity for NMDA receptors, was found to have a rapid and robust antidepressant effect within 2 h in several large randomised controlled trials (RCT) of people with MDD (Singh et al. , 2016), and has now been approved within the United States for intranasal administration for people with high risk of suicide (Henter et al., 2018).

Two subunit NR2B-specific NMDA receptor antagonists were recently tested for MDD. While CP-101,606 (traxoprodil) was effective but was halted due to cardiovascular toxicity, MK-0657 (CERC-301) had no significant side effects but had mixed outcomes (Henter et al., 2018). Rapastinel, a glycine partial NMDA agonist, has shown high efficacy in clinical trials for major depression disorder (MDD), and has now been approved for the adjunctive treatment of MDD in the United States (Moskal et al., 2014; Preskorn et al., 2015; Vasilescu et al., 2017). Preliminary results show that sarcosine, a glycine transporter-I inhibitor that potentiates NMDA function, was more effective that citalopram, with no significant side effects (Huang et al., 2013). 4-Cl-KYN (AV-101), a highly selective glycine receptor antagonist, was highly effective in animal studies and is now being tested in clinical trials for MDD (Zanos et al. , 2015). Additionally, there are agents that target the mGluRs, but none have been demonstrated to achieve a strong anti-depressive effect (Henter et al., 2018). Thus, the mechanisms and effectiveness of several glutaminergic agents require further study.

Natural Boosts for Everyday Functioning

Another reason to get glutamate into the public eye is that with minimal knowledge of its mechanisms, there are many natural ways the lay public can boost their overall health and wellbeing. Physical exercise and mindfulness exercises have both been demonstrated to be powerful modulators of non-pharmaceutical glutamate and GABA interventions.

Physical exercise leads to increase levels of both glutamate and GABA (Maddock et al., 2016), resulting in participants feeling energized and focused while also experiencing psychological calm. In adult rats, running has been demonstrated to stimulate neurogenesis and increase the gene expression levels of the NR2B subunit of the NDMA receptor in the dentate gyrus, leading to enhanced learning, memory, and mood functioning (Vivar and van Praag, 2017). In humans, three different experiments show that vigorous physical activity results increased content of glutamate and GABA in the visual and anterior cingulate cortices in comparison with sedentary activity (Maddock et al., 2016). Levels rose approximately five percent and persisted for at least 30 min post-exercise. Additionally, participants who had higher levels of exercise in the previous week also had higher resting glutamate levels.

Mindfulness has a strong impact on brain glutamate levels observed in the brains of people who meditate mindfulness (Fayed et al., 2013). A cross-sectional study comparing the brains of meditators from a Zen Buddhist monastery with hospital staff showed a negative correlation between years of meditation and levels of glutamate in the left thalamus, which may indicate a higher level of efficiency of glutamate metabolism in this area (Fayed et al., 2013). The Zen meditators also had high myo-inositol concentrations in the posterior cingulate, which may indicate higher levels of glial and microglial activation. The exact mechanisms by which glutamate may modulate the effects of mindfulness still must be explored.

Conclusion

This brief review has highlighted the widespread impact of glutamate throughout the brain health. Glutamate is critical for maintenance of ideal energy levels, necessary for most CNS functions, and neuroplasticity, which is critical for adaptation to changes in the environment. Rather than being delegated as a sidenote as in the past, glutamate is deserving of a main focus in future neuroscience research and clinical studies. Additionally, efforts should be made to educate the lay public as to the importance of glutamate to everyday functioning and how to maintain healthy levels for increased resiliency in times of stress.

Author Contributions

The author confirms being the sole contributor of this work and has approved it for publication.

Conflict of Interest

MP is the Director at In Cognition UK, private clinic and sole author to the presented research. This research has been conducted in the absence of any commercial or financial relationship that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Arnone, D., Mumuni, A. N., Jauhar, S., Condon, B., and Cavanagh, J. (2015). Indirect evidence of selective glial involvement in glutamate-based mechanisms of mood regulation in depression: meta-analysis of absolute prefrontal neuro-metabolic concentrations. Eur. Neuropsychopharmacol. 25, 1109–1117. doi: 10.1016/j.euroneuro.2015.04.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Baker, D. A., Xi, Z. X., Shen, H., Swanson, C. J., and Kalivas, P. W. (2002). The origin and neuronal function of in vivo nonsynaptic glutamate. J. Neurosci. 22, 9134–9141. doi: 10.1523/jneurosci.22-20-09134.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

Chitty, K. M., Lagopoulos, J., Lee, R. S., Hickie, I. B., and Hermens, D. F. (2013). A systematic review and meta-analysis of proton magnetic resonance spectroscopy and mismatch negativity in bipolar disorder. Eur. Neuropsychopharmacol. 23, 1348–1363. doi: 10.1016/j.euroneuro.2013.07.007

PubMed Abstract | CrossRef Full Text | Google Scholar

de Sousa, R. T., Loch, A. A., Carvalho, A. F., Brunoni, A. R., Haddad, M. R., and Henter, I. D. (2017). Genetic studies on the tripartite glutamate synapse in the pathophysiology and therapeutics of mood disorders. Neuropsychopharmacology 42, 787–800. doi: 10.1038/npp.2016.149

PubMed Abstract | CrossRef Full Text | Google Scholar

Fayed, N., Lopez Del Hoyo, Y., Andres, E., Serrano-Blanco, A. , Bellón, J., Aguilar, K., et al. (2013). Brain changes in long-term zen meditators using proton magnetic resonance spectroscopy and diffusion tensor imaging: a controlled study. PLoS One 8:e58476. doi: 10.1371/journal.pone.0058476

PubMed Abstract | CrossRef Full Text | Google Scholar

Gigante, A. D., Bond, D. J., Lafer, B., Lam, R. W., Young, L. T., and Yatham, L. N. (2012). Brain glutamate levels measured by magnetic resonance spectroscopy in patients with bipolar disorder: a meta-analysis. Bipolar Disord. 14, 478–487. doi: 10.1111/j.1399-5618.2012.01033.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, X., Rao, Y., Mao, R., Cui, L., and Fang, Y. (2020). Common cellular and molecular mechanisms and interactions between microglial activation and aberrant neuroplasticity in depression. Neuropharmacology 181:108336. doi: 10.1016/j.neuropharm.2020.108336

PubMed Abstract | CrossRef Full Text | Google Scholar

Henter, I. D., de Sousa, R. T., and Zarate, C. A. Jr. (2018). Glutamatergic modulators in depression. Harv. Rev. Psychiatry 26, 307–319. doi: 10.1097/HRP.0000000000000183

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, C. C., Wei, I. H., Huang, C. L., Chen, K. T., Tsai, M. H., and Tsai, P. (2013). Inhibition of glycine transporter-I as a novel mechanism for the treatment of depression. Biol. Psychiatry 74, 734–741. doi: 10.1016/j.biopsych.2013.02.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Kantrowitz, J. T., Halberstam, B., and Gangwisch, J. (2015). Single-dose ketamine followed by daily D-Cycloserine in treatment-resistant bipolar depression. J. Clin. Psychiatry 76, 737–738. doi: 10.4088/JCP.14l09527

PubMed Abstract | CrossRef Full Text | Google Scholar

Lynch, M. A. (2004). Long-term potentiation and memory. Physiol. Rev. 84, 87–136.

Google Scholar

Maddock, R. J., Casazza, G. A., Fernandez, D. H., and Maddock, M. I. (2016). Acute modulation of cortical glutamate and GABA content by physical activity. J. Neurosci. 36, 2449–2457. doi: 10.1523/JNEUROSCI.3455-15.2016

PubMed Abstract | CrossRef Full Text | Google Scholar

Mandal, S., Sinha, V. K., and Goyal, N. (2019). Efficacity of ketamine therapy in the treatment of depression. Indian J. Psychiatry 61, 480–485. doi: 10.103/psychiatry.IndianJPsychiatry_484_18

CrossRef Full Text | Google Scholar

Moskal, J. R., Burch, R., Burgdorf, J. S., Kroes, R. A., Stanton, P. K., Disterhoft, J. F., et al. (2014). GLYX-13, an NMDA receptor glycine site functional partial agonist enhances cognition and produces antidepressant effects without the psychotomimetic side effects of NMDA receptor antagonists. Expert Opin. Investig. Drugs 23, 243–254. doi: 10.1517/13543784.2014.852536

PubMed Abstract | CrossRef Full Text | Google Scholar

Newport, D. J., Carpenter, L. L., McDonald, W. M., Potash, J. B., Tohen, M., and Nemeroff, C. B. (2015). Ketamine and other NMDA antagonists: early clinical trials and possible mechanisms in depression. Am. J. Psychiatry 172, 950–966. doi: 10.1176/appi.ajp.2015.15040465

PubMed Abstract | CrossRef Full Text | Google Scholar

Preskorn, S., Macaluso, M., Mehra, D. O., Zammit, G., Moskal, J. R., Burch, R. M., et al. (2015). Randomized proof of concept trial of GLYX-13, an N-methyl-D-aspartate receptor glycine site partial agonist, in major depressive disorder nonresponsive to a previous antidepressant agent. J. Psychiatr. Pract. 21, 140–149. doi: 10.1097/01.pra.0000462606.17725.93

CrossRef Full Text | Google Scholar

Reznikov, L. R., Fadel, J. R., and Reagan, L. P. (2011). “Glutamate-mediated neuroplasticity deficits in mood disorders,” in Neuroplasticity, eds J. A. Costa e Silva, J. P. Macher, and J. P. Olié (Tarporley: Springer), 13–26. doi: 10.1007/978-1-908517-18-0_2

CrossRef Full Text | Google Scholar

Sah, P., Westbrook, R. F., and Luthi, A. (2008). Fear conditioning and long-term potentiation in the amygdala: what really is the connection? Ann. N. Y. Acad. Sci. 1129, 88–95. doi: 10.1196/annals.1417.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Singh, J. B., Fedgchin, M., Daly, E., Xi, L., Melman, C., and De Bruecker, G. (2016). Intravenous esketamine in adult treatment-resistant depression: a double-blind, double-randomization, placebo-controlled study. Biol. Psychiatry 80, 424–431. doi: 10.1016/j.biopsych.2015.10.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Vasilescu, A. N., Schweinfurth, N., Borgwardt, S., Gass, P., Lang, U. E., Inta, D., et al. (2017). Modulation of the activity of N-methyl-d-aspartate receptors as a novel treatment option for depression: current clinical evidence and therapeutic potential of rapastinel (GLYX-13). Neuropsychiatr. Dis. Treat. 13, 973–980. doi: 10.2147/NDT.S119004

PubMed Abstract | CrossRef Full Text | Google Scholar

Vivar, C. , and van Praag, H. (2017). Running changes the brain: the long and the short of it. Physiology 32, 410–424. doi: 10.1152/physiol.00017.2017

PubMed Abstract | CrossRef Full Text | Google Scholar

Watkins, J. C., and Jane, D. E. (2006). The glutamate story. Br. J. Pharmacol. 147, 100–108. doi: 10.1038/sj.bjp.0706444

PubMed Abstract | CrossRef Full Text | Google Scholar

Zanos, P., Moaddel, R., Morris, P. J., Georgiou, P., Fischell, J., and Elmer, G. I. (2016). NMDAR inhibition-independent antidepressant actions of ketamine metabolites. Nature 533, 481–486. doi: 10.1038/nature17998

PubMed Abstract | CrossRef Full Text | Google Scholar

Zanos, P., Piantadosi, S. C., Wu, H. Q., Pribut, H. J., Dell, M. J., and Can, A. (2015). The prodrug 4-chlorokynurenine causes ketamine-like antidepressant effects, but not side effects, by NMDA/glycineB-site inhibition. J. Pharmacol. Exp. Ther. 355, 76–85. doi: 10.1124/jpet.115.225664

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhou, Y. , and Danbolt, N. C. (2014). Glutamate as a neurotransmitter in the healthy brain. J. Neural Transm. 121, 799–817. doi: 10.1007/s00702-014-1180-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Glutamate – Disease & Food

By Lindsey KonkelMedically Reviewed by Rosalyn Carson-DeWitt, MD

Reviewed:

Medically Reviewed

The neurotransmitter glutamate is produced in your body, and is also found in many foods.

Glutamate is a neurotransmitter that sends signals in the brain and throughout the nerves in the body.

Glutamate plays an important role during brain development. Normal levels of glutamate also help with learning and memory.

Having too much glutamate in the brain has been associated with neurological diseases such as Parkinson’s disease, multiple sclerosis, Alzheimer’s disease, stroke, and ALS (amyotrophic lateral sclerosis or Lou Gehrig’s disease).

Problems in making or using glutamate have also been linked to a number of mental health disorders, including autism, schizophrenia, depression, and obsessive-compulsive disorder (OCD).

Glutamate and Disease

Glutamate has many important functions in the brain, in addition to passing chemical messages from one nerve cell to another.

Too much glutamate may damage nerve cells and the brain.

There are two ways that glutamate can be damaging: There can be too much glutamate in the brain, or the receptors for glutamate on receiving nerve cells may be oversensitive, meaning that fewer glutamate molecules are needed to excite them.

At high concentrations, glutamate can overexcite nerve cells, causing them to die. Prolonged excitation is toxic to nerve cells, causing damage over time. This is known as excitotoxicity.

Researchers are studying therapies that attempt to inhibit glutamate activity for the treatment of ALS.

Glutamate and Food

Glutamate is a naturally occurring amino acid found in many different types of food. Amino acids are the building blocks of protein.

Glutamate is perhaps best known as the food additive monosodium glutamate (MSG).

MSG is used as a flavor enhancer commonly found in American-style Chinese food, canned soups and vegetables, and processed meats.

MSG can also be found naturally in many foods, including tomatoes, cheeses, mushrooms, seaweed, and soy.

While some people report adverse reactions to MSG, such as headaches, nausea, or heart palpitations, researchers have found no definitive link between MSG and these symptoms.

According to the Food and Drug Administration (FDA), MSG is generally safe at the levels commonly found in the typical American diet.

By subscribing you agree to the Terms of Use and Privacy Policy.

Editorial Sources and Fact-Checking

  • Glutamate brain basics, National Institute of Mental Health
  • Glutamate and food, U.S. FDA
  • Glutamate and disease, The ALS Association
  • Glutamate and excitotoxicity, Stanford University

Show Less

What Is Norepinephrine?

Norepinephrine is a natural chemical in the body that’s released by stress during the fight-or-flight response. It also affects mood and attention.

By Cathy Cassata

Melatonin Side Effects and Safety 101

Before you take melatonin for better sleep, make sure you understand the potential side effects.

By Valencia Higuera

7 Melatonin Mistakes Sleep Doctors Want You to Avoid

Over-the-counter melatonin supplements are widely used and available. Doctors say both the safety and effectiveness of this supplement depend on how it…

By Lisa Rapaport

Hormones and Your Health: An Essential Guide

Hormones are vital chemicals that enable daily bodily functions, reproduction, movement, and more. Learn about cortisol and stress; serotonin, dopamine…

By Lindsey Konkel

What Is Mucus? Symptoms, Causes, Diagnosis, Treatment, and Prevention

Mucus — also known as sputum — is a sticky, gelatinous material that lines your lungs, throat, mouth, nose, and sinuses.

By Brian P. Dunleavy

What Is Progesterone?

Sometimes called the ”pregnancy hormone,” progesterone is a hormone produced by a woman’s ovaries. The corpus luteum, a temporary endocrine gland, secretes…

By Cathy Cassata

Cortisol: The Stress Hormone

Cortisol is a steroid hormone that helps the body respond to stress. As part of the body’s fight-or-flight response, cortisol is released during stressful…

By Lindsey Konkel

What Is Dopamine?

Dopamine is a hormone involved mainly in controlling movement, but it also plays a role in the brain’s reward system, helping to reinforce certain behaviors…

By Lindsey Konkel

What Is Serotonin?

Serotonin, a hormone and neurotransmitter, is sometimes known as the happy chemical. It appears to play a role in regulating mood. Low levels of serotonin…

By Lindsey Konkel

Parkinson’s disease and the glutamatergic system

Parkinson’s disease and the glutamatergic system

Website of the publishing house “Media Sfera”
contains materials intended exclusively for healthcare professionals. By closing this message, you confirm that you are a registered medical professional or student of a medical educational institution.

Mironova Yu.S.

Siberian State Medical University, Tomsk, Russia

Zhukova N.G.

Siberian State Medical University of the Ministry of Health of Russia, Tomsk, Russia

Zhukova I.A.

Department of Neurology and Neurosurgery, Siberian State Medical University of the Ministry of Health of the Russian Federation, Tomsk

Alifirova V.M.

Siberian State Medical University, Tomsk

Izhboldina O. P.

Department of Neurology and Neurosurgery, Siberian State Medical University of the Ministry of Health of the Russian Federation, Tomsk

Latypova A.V.

Siberian State Medical University, Tomsk, Russia

Parkinson’s disease and glutamatergic system

Authors:

Mironova Yu.S., Zhukova N.G., Zhukova I.A., Alifirova V.M. , Izhboldina O.P., Latypova A.V.

More about the authors

Magazine:

Journal of Neurology and Psychiatry. S.S. Korsakov.

2018;118(5): 138-142

DOI:

10.17116/jnevro201811851138

How to quote:

Mironova Yu.S., Zhukova N.G., Zhukova I.A., Alifirova V.M., Izhboldina O.P., Latypova A.V. Parkinson’s disease and the glutamatergic system. Journal of Neurology and Psychiatry. S.S. Korsakov.
2018;118(5):138-142.
Mironova YuS, Zhukova NG, Zhukova IA, Alifirova VM, Izhboldina OP, Latypova AV. Parkinson’s disease and glutamatergic system. Zhurnal Nevrologii i Psikhiatrii imeni S. S. Korsakova. 2018;118(5):138-142. (In Russ.)
https://doi.org/10.17116/jnevro201811851138

Read metadata

Clinical variants of Parkinson’s disease (PD) are not limited to a complex of motor manifestations, but include a wide range of different non-motor symptoms: cognitive, psychotic, autonomic and sensory. Often, it is non-motor manifestations that precede the development of the classical motor picture of the disease, which has recently been actively studied in the course of neuroimaging, pathomorphological and genetic studies. Therefore, today PD is considered as a multisystem, heterogeneous disorder associated with multineurotransmitter dysfunction. This leads to the understanding that not only dopaminergic, but also other neurotransmitter systems, including glutamatergic, are involved in the pathogenesis of PD. The article deals with the participation of the glutamatergic system in the formation of the neurodegenerative process. The role of glutamate as a neurotransmitter and neurotoxin in the pathogenesis of PD and in the development of its clinical manifestations is discussed. It is assumed that the study of the state of glutamate excitotoxicity in patients with PD will improve treatment tactics and correct pathogenetic therapy.

Keywords:

Parkinson’s disease

glutamate

excitotoxicity

Authors:

Mironova Yu.S.

Siberian State Medical University, Tomsk, Russia

Zhukova N.G.

Siberian State Medical University of the Ministry of Health of Russia, Tomsk, Russia

Zhukova I.A.

Department of Neurology and Neurosurgery, Siberian State Medical University of the Ministry of Health of the Russian Federation, Tomsk

Alifirova V.M.

Siberian State Medical University, Tomsk

Izhboldina O.P.

Department of Neurology and Neurosurgery, Siberian State Medical University of the Ministry of Health of the Russian Federation, Tomsk

Latypova A.V.

Siberian State Medical University, Tomsk, Russia

Close metadata

Currently, the issues of the occurrence of Parkinson’s disease (PD), the role of possible triggering factors in the development of early non-motor and basic motor disorders, as well as the causes of certain clinical variants of the course of the disease, are being actively studied. It is generally accepted that the key mechanism for the development of PD is the degeneration of dopaminergic neurons in the substantia nigra. However, in recent years, researchers have been paying more and more attention to the involvement of other neurotransmitter systems in the processes of pathogenesis and the formation of various clinical manifestations of PD [1, 2].

There are several hypotheses regarding the causes of the death of dopamine-producing neurons of the substantia nigra in PD, including genetic predisposition, oxidative stress, impaired mitochondrial respiration, the action of the neurotoxin glutamate, leading to increased activation of postsynaptic glutamate receptors, and to changes in the permeability of ion channels, excessive supply of ions Ca 2+ into cells, which increases the activity of intracellular proteases [3]. According to modern concepts, all of the above mechanisms can participate in the pathogenesis of neurodegeneration in PD simultaneously or sequentially potentiating each other. In this regard, PD can be considered as a multifactorial disease that manifests itself as a result of the interaction of genetic and environmental factors [4]. This interaction ultimately leads to the activation of apoptosis, but the trigger mechanism, as well as the sequence of pathogenetic factors of neurodegeneration, are still unclear [5].

In recent years, along with impaired dopaminergic neurotransmission, more and more evidence has been found of the important role of excitotoxicity in the development of PD [6, 7]. Excitotoxicity leads to damage and death of nerve cells due to the action of neurotransmitters that activate specific postsynaptic NMDA (N-methyl-D-aspartic acid) and AMPA (alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid) receptors , and induction of apoptosis processes. An important role in this process is played by excitatory neurotransmitters, primarily glutamate [8].

Glutamate is one of the most common excitatory neurotransmitters in the vertebrate nervous system [9]. Glutamic acid itself was discovered and described by the German chemist K.Kh. Ritthausen in 1866 during the treatment of wheat gluten (gluten) with sulfuric acid [10]. Later, in 1908, the Japanese researcher at the Tokyo Imperial University K. Ikeda discovered brown crystals left after the evaporation of a large amount of broth formed by boiling edible kelp algae, which turned out to be glutamic acid. Subsequently, this method of mass production of a crystalline salt of glutamic acid (monosodium glutamate) was patented by him [11].

For the first time the idea of ​​the importance of glutamate, or rather its negative effect on the nervous system, was expressed by the Japanese researcher C. Hayashi in 1954. In his works, the introduction of glutamic acid into the ventricles of the brain of dogs and monkeys caused convulsions in animals. Towards the end of the 1950s, experiments were carried out in which glutamate depolarized and excited spinal cord neurons in cats. D. Lucas and J. Newhouse in 1957 [12] revealed the death of retinal neurons when mice were injected with monosodium glutamate. At 19J. Olney [13] discovered that this phenomenon is not limited to the retina, but affects all structures of the CNS, and proposed the term “glutamate excitotoxicity”. The scientist identified a correlation between the excitatory properties and neurotoxicity of glutamate, and also established the ability of glutamate antagonists to block neurotoxicity. Subsequently, these results were reproduced in a primate model using higher doses of glutamate [14]. These studies have shown that the hypothalamus and periventricular areas of the brain are particularly sensitive to glutamate. A similar neuroanatomical picture of degeneration was observed after brain hypoxia [15], which suggested a possible role of glutamate excitotoxicity in the process of ischemic neuronal death. Evidence supporting this theory was provided by S. Rothman [16], who demonstrated reduced sensitivity to hypoxia in hippocampal cell cultures using γ-D-glutamylglycine (an inotropic glutamate antagonist), a nonspecific inhibitor of the postsynaptic excitatory amino acid. Subsequent studies by M. Mattson et al. [17] showed the role of glutamate excitotoxicity in Alzheimer’s disease, PD, and other neurodegenerative diseases associated with oxidative stress and cellular energy deficiency. In their early studies, they found that the neurotransmitter glutamate, previously thought to function only in synapses, plays a key role in the regulation of dendritic development and synaptogenesis [18].

In neurons, glutamate is initially synthesized from the carbon of oxidized glucose through the formation of ketoglutarate in the reaction of the tricarboxylic acid cycle [19]. The mechanism by which glutamate exits the synaptic cleft is not fully understood. It is known that the release of glutamate from the synaptic vesicle occurs through exocytosis [20] from the cytosol through activated anion channels with the participation of mediator proteins [21]. Glutamate is released into the synaptic cleft and then enters astrocytes [22], where it is aminated by glutamine synthase to glutamine. The latter enters the presynaptic endings of glutamatergic neurons, turning into glutamate with the help of the phosphate-activated form of the glutaminase enzyme. The brain-specific transporter is localized selectively at the terminals of glutamatergic neurons and thus can regulate the replenishment of the glutamate pool in presynaptic terminals [19].]. It has been shown that any disturbance in the ability of astrocytes to maintain an optimal extracellular level of glutamate by functioning transport systems is closely associated with neurodegeneration due to excitotoxicity [23–25].

An important role as the main regulator of glutamate transport and its extracellular concentration is played by special transporters – glutamate transporters ( eng .: excitatory amino acid transporter – EAAT). Five Na 9 have been identified to date0096 + -dependent glutamate transporters, which are located on various brain cells [26]. It is known that EAAT1, 2, and 3 are found in endothelial cells and astrocytes, while EAAT3 is found in neurons. This indicates that glutamate is constantly circulating between endothelial cells, astrocytes, and neurons. The main function of EAAT1 and EAAT2 is considered to be the removal of glutamate from the synaptic cleft for the purpose of its subsequent conversion into glutamine. It turned out that their work depends on the concentration of extracellular glutamate [27]. It becomes obvious that glutamate itself and its transporters are involved both in the normal functioning of the brain and in the development of the neurodegenerative process. Thus, one of the key roles in the development of excitotoxicity in PD can be played by the dysfunction of glutamate transporters [28].

It is known that glutamate is one of the main excitatory mediators of the CNS, however, excessive activation of its receptors can lead to processes leading to neuronal death [29]. There are two subtypes of glutamate receptors, based on their role in the functioning of nerve cells. The first group includes ionotropic ones, which are structurally associated with ion channels and open only after they are activated by special ligands in such a way that the electrical activity of the neuron is caused by ion flows. The second group includes metabotropic receptors that are not associated with ion channels, which, using special signaling molecules, regulate metabolic processes in cells and control the activity of ionotropic receptors [30]. Metabotropic receptors are characterized by both post- and presynaptic localization and expression not only in neuronal but also in glial cells. These receptors are involved in the implementation of the mechanisms of memory, pain perception, anxiety states, as well as in the processes of neurodegeneration [31, 32]. To date, it is customary to distinguish 3 main subtypes of ionotropic glutamate receptors based on sensitivity to the most selective agonist: sensitive to NMDA, to AMPA and preferably responsive to kainic acid – kainate receptors.

Much attention is paid to NMDA receptors, since it is with their function that the increase in synaptic transmission between two neurons is associated, which persists for a long period of time after exposure to the synapse, excitation (long-term potentiation) in the hippocampus, the ability to perceive damaging actions through nociceptors (nociception), the formation electrical activity of the brain in the form of sharp waves or peaks that differ from background activity (epileptiform activity), as well as excitotoxic effects of glutamate [33]. The receptor is a heteromeric complex that interacts with several intracellular proteins with three different subunits: NMDAR 1 (NR1), NMDAR 2 (NR2), and NMDAR 3 (NR3) [34]. The NR1 subunit has 8 different variants generated by the process of synthesis of multiple isoforms of receptor subunits, the so-called alternative splicing, and encoded by one gene – GRIN1 ( EN : glutamate ionotropic receptor NMDA type subunit 1). It has been established that NR2 exists in 4 varieties (A, B, C and D), each of which is encoded by a separate gene, respectively GRIN2A , GRIN2B , GRIN2C , GRIN2D . The involvement of the NR2B subunit in the processes of learning and memory, as well as in the formation of chronic pain syndrome [35], has been established. At the end of the 20th century, NR3A and NR3B were discovered [36], which are currently being studied.

An important difference between NMDA glutamate receptors and other ionotropic receptors is that their channel transmits not only Na + and K + ions, but also Ca 2+ and is a second messenger and is able to modulate the cell response depending on external signal [37]. The highest density of NMDA receptors was found in the hippocampus, cerebral cortex, amygdala, and striatum [38]. This is probably why, under normal conditions, the activation of NMDA receptors is associated with the plasticity of CNS structures, learning and memory processes [39]. At present, a significant contribution of hyperactivation of ionotropic NMDA receptors to the process of death of dopaminergic neurons of the substantia nigra has been established. The mechanism of protection of neurons in the substantia nigra from the toxic effects of methamphetamine due to blockade of NMDA receptors has been shown [40, 41]. To date, there is evidence of the ability of glutamate in millimolar concentrations, which is in the synaptic cleft for several milliseconds, to activate NMDA receptors under physiological conditions [42, 43]. However, in pathology, these same receptors can be activated by lower, micromolar concentrations of glutamate, but for a significantly longer time [44, 45]. Hyperactivation of ionotropic glutamate receptors leads to a sharp increase in the transmembrane calcium current into the cell, followed by the release of Ca 2+ from intracellular depots, depolarization of the mitochondrial membrane and, as a result, a long-term increase in the amount of Ca 2+ in the cytoplasm. The high content of Ca 2+ in neurons triggers neurotoxic processes with the activation of proteolytic enzymes and the destruction of cellular structures, which ultimately leads to increased synthesis of nitric oxide, activation of lipid peroxidation and, as a result, to oxidative stress, disruption of the synthesis of neurotrophic factors, as well as to apoptosis [46, 47]. It follows from this that glutamate excitotoxicity can not only trigger but also aggravate the neurodegenerative process in PD.

The direct role of endogenous glutamate in the death of neurons in the substantia nigra has also been shown in ischemic brain damage, in which there is a significant increase in the concentration of extracellular glutamate [48, 49]. However, in addition to the process of acute (“classical”) excitotoxicity, it is customary to isolate the mechanism of the so-called slow (“metabolic”) excitotoxicity, which develops in neurodegenerative diseases [50].

The formation of neurodegenerative diseases such as PD is possible with changes in the functional state of receptors, metabolic disorders, and neuron energy deficiency [51, 52]. Normally, the defense systems of the brain are able to eliminate the neurotoxic effect of glutamate, since NMDA receptors are blocked by extracellular Mg ions 2+ , however, this blockade is a potential-dependent process, and this mechanism is turned off when neurons depolarize. The function of maintaining membrane polarization is mainly provided by ion pumps, in particular Na + /K + -adenosine triphosphatase (Na + /K + -ATPase) [53], and the transport of ions into and out of cells is an energy-dependent process . The delayed onset and slow progression of neurodegenerative diseases is based on a genetically determined defect in energy metabolism, which realizes its damaging effect only after activation of normal aging processes [54]. The decrease in the concentration of ATP in neurons, observed during aging, leads to a decrease in the activity of ATPase, which leads to membrane depolarization. It follows that even the natural concentration of extracellular glutamate can be toxic [55]. In studies in vitro on neuronal cultures, inhibition of oxidative phosphorylation by cyanides or inactivation of Na + /K + -ATPase by ouabain (strophanthin G, the most potent member of the cardiac glycoside group) led to the fact that low concentrations of glutamate became fatal, in both cases blockade of NMDA receptors prevented cell death [56].

To date, data obtained under experimental conditions indicate the likely involvement of glutamate excitotoxicity in neuronal degeneration of the Alzheimer’s type [57]. As confirmation, the fact that glutamate, as a fast neurotransmitter in brain structures, is actively involved in the functioning of the hippocampus and the cerebral cortex due to the regulation of learning and memory mechanisms can serve as confirmation [50, 58]. However, under certain conditions, it can also become an excitotoxin, taking a direct part in the neurodegenerative process [29].].

It is known that glutamate is considered as the predominant excitatory neurotransmitter in the basal ganglia, which are responsible for motor function, and in their pathology there is a motor deficit pathognomonic for PD [59]. Earlier studies in experimental animal models of PD revealed dysregulation of glutamatergic systems [60–62]. Striatal dopamine deficiency induced by injury to the nigrostriatal tract using 6-OHDA (6-hydroxydopamine—a toxin for selective killing of dopaminergic neurons in the brain) has also been shown to result in increased glutamate release [63].

Currently, the understanding of the role of glutamate as a neurotransmitter and neurotoxin in the pathogenesis of PD and the development of its clinical manifestations remains insufficiently elucidated. In this regard, an in-depth study of glutamate excitotoxicity is important in order to determine early diagnostic biomarkers of the pathological process in PD, as well as to search for new therapeutic approaches in the treatment of this disease, when neurodegeneration processes have affected only a limited number of cells.

The authors declare no conflict of interest.

*e-mail: [email protected]



Serum glutamate in patients with schizophrenia spectrum disorders and bipolar affective disorder | Seregin

1. Ohgi Y, Futamura T, Hashimoto K. Glutamate signaling in synaptogenesis and NMDA receptors as potential therapeutic targets for psychiatric disorders. Curr. Mol. Med. 2015;15:206–221. https://DOI:10.2174/1566524015666150330143008

2. Li CT, Lu CF, Lin HC, Huang YZ, Juan CH , Su TP, Bai YM, Chen MH, Lin WC. Cortical inhibitory and excitatory function in drug-naive generalized anxiet y disorder. Brain Stimulus. 2017;10(3):604–608.https://DOI:10.1016/j.brs.2016.12.007

3. Krystal JH, Karper LP, Seibyl JP, Freeman GK, Delaney R, Bremner JD, Heninger GR , Bowers MB Jr, Charney DS. Subanesthetic effects of the noncompetitive NMDA antagonist, ketamine, in humans. Psychotomimetic, perceptual, cognitive, and neuroendocrine responses. Arch. Gen. Psychiatry. 1994;51:199–214. DOI:10.1001/archpsyc.1994.03950030035004

4. Javitt DC. Negative schizophrenic symptomatology and the PCP (phencyclidine) model of Hillside schizophrenia. J.Clin. Psychiatry. 1987;(9):12–35.

5. Javitt DC, Zukin SR, Heresco-Levyt U, Umbricht D. Has an angel shown the way? Etiological and therapeutic implications of the PCP/NMDA model of schizophrenia. Schizophr. Bull. 2012;(38):958–966. https://DOI:10.1093/schbul/sbs069

6. Burbaeva GSH, Boksha IS, Starodubtseva LI, Savushkina OK, Tereshkina EB, Turishcheva MS, Prokhorova TA, Vorob’eva EA, Morozova MA. Violation of glutamate metabolism in schizophrenia. Bulletin of the Russian Academy of Medical Sciences. 2007;3:19-24.

7. Toru M, Kurumaji A, Ishimaru M. Excitatory amino acids: implications for psychiatric disorders research. life sci. 1994;(55):1683–1699.

8. Kerwin R, Patel S, Meldrum B. Quantitative autoradiographic analysis of glutamate binding sites in the hippocampal formation in normal and schizophrenic brain post mortem. neuroscience. 1990;(39):25–32.

9. Beneyto M, Meador-Woodruff JH. Lamina-specific abnormalities of NMDA receptor-associated postsynaptic protein transcripts in the prefrontal cortex in schizophrenia and bipolar disorder. neuropsychopharmacology. 2008;(33):2175–2186. https://DOI:10.1038/sj.npp.1301604

10. Beneyto M, Kristiansen LV, Oni-Orisan A, McCullumsmith RE, Meador-Woodruff JH. Abnormal glutamate receptor expression in the medial temporal lobe in schizophrenia and mood disorders. neuropsychopharmacology. 2007;(3 2):1888–1902. https://doi:10.1038/sj.npp.1301312

11. Matosin N, Fernandez-Enright F, Frank E, Deng C, Wong J, Huang XF, Newell KA. Metabotropic glutamate receptor mGluR2/3 and mGluR5 binding in the anterior cingulate cortex in psychotic and nonpsychotic depression, bipolar disorder and schizophrenia: implications for novel mGluR-based therapeutics. J. Psychiatry Neurosci. 2014;(39):407–416. https://DOI:10.1503/jpn.130242

12. Blacker CJ, Lewis CP, Frye MA, Veldic M. Metabotropic glutamate receptors as emerging research targets in bipolar disorder. Psychiatry Res. 2017;(257):327–337. https://doi:10.1016/j.psychres.2017.07.059

13. Gigante AD, Bond DJ, Lafer B, Lam RW, Young LT, Yatham LN. Brain glutamate levels measured by magnetic resonance spectroscopy in patients with bipolar disorder: a meta-analysis. Bipolar Discord. 2012;(14):478–487. https://DOI:10.1111/j.1399-5618.2012.01033.x

14. Chitty KM, Lagopoulos J, Lee RS, Hickie IB, Hermens DF. A systema tic review and meta-analysis of proton magnetic resonance spectroscopy and mismatch negativity in bipolar disorder. Eur. Neuropsychopharmacol. 2013;(23):1348–1363. https://DOI:10.1016/j.euroneuro.2013.07.007

15. Loginova LV, Smirnova LP, Seregin AA, Dmitrieva EM, Mazin EV, Simutkin GG. To the question of the search for biomarkers in bipolar affective disorder. Bulletin of the Ural medical academic science. 2014;3(49):139–141.

16. Merritt K, Egerton A, Kempton MJ, Taylor MJ, McGuire PK. Nature of glutam ate alterations in schizophrenia: a meta-analysis of proton magnetic resonance spectroscopy studies. JAMA Psychiatry. 2016;(73):665–674. https://doi:10.1001/jamapsychiatry.2016.0442

17. Egerton A, Broberg BV, Van Haren N, Merritt K, Barker GJ, Lythgoe DJ, Perez-Iglesias R, Baandrup L, Düring SW, Sendt KV, Stone JM , Rostrup E, Sommer IE, Glenthøj B, Kahn RS, Dazzan P, McGuire P. Response to initial antipsychotic treatment in first episode psychosis is related to anterior cingulate glutamate levels: a multicentre (1)H-MRS study (OPTiMiSE). Mol. Psychiatry. 2018;(23):2145–2155. https://DOI:10.1038/s41380-018-0082-9.

18. Nugent AC, Diazgranados N, Carlson PJ, Ibrahim L, Luckenbaugh DA, Brutsche N, Herscovitch P, Drevets WC, Zarate CA Jr. Neural correlates of rapid antidepressant response to ketamine in bipolar disorder. Bipolar Discord. 2014;(16):119–128. https://DOI:10.1111/bdi.12118

19. Altamura CA, Mauri MC, Ferrara A, Moro AR, D’Andrea G, Zamberlan F. Plasma and platelet excitatory amino acids in psychiatric disorders. Am. J. Psychiatry. 1993;150(11):1731–1733.

20. Semke AV, Vetlugina TP, Ivanova SA, Rakhmazova LD, Gutkevich EV, Lobacheva OA, Kornetova EG. Biopsychosocial foundations and adaptive-compensatory mechanisms of schizophrenia in the Siberian region. Siberian Bulletin of Psychiatry and Narcology. 2009;5(56):15–20.

21. Semke AV, Fedorenko OY, Lobacheva OA, Rakhmazova LD, Kornetova EG, Smirnova LP, Mikilev FF, Shchigoreva YuG. Clinical, epidemiological and biological prerequisites for the adaptation of patients with schizophrenia as the basis for a personalized approach to antipsychotic therapy. Siberian Bulletin of Psychiatry and Narcology. 2015;3(88):19–25.

22. Inoshita M, Umehara H, Watanabe SY, Nakataki M, Kinoshita M, Tomioka Y, Tajima A, Numata S, Ohmori T. Elevated peripheral bl ood glutamate levels in major depressive disorder. Neuropsych. Dis. & treat. 2018;(14):945–953. https://DOI:10.2147/NDT.S159855

23. Smirnova LP, Loginova LV, Ivanova SA, Dmitrieva EM, Seregin AA, Mikilev FF, Semke AV, Bokhan NA. Laboratory method for the diagnosis of schizotypal disorder. Pat. No. 2569741 Russian Federation G01N 33/50. 2014148200/15; publ. 11/27/2015.

24. Purcell SM, Wray NR, Stone JL, Visscher PM, O’Donovan MC, Sullivan PF, Sklar P. Common polygenic variation contributes to risk of schizophrenia and bipolar disorder. Nature. 2009;460(7256):748–752. https://DOI:10.1038/nature08185

25. Bipolar Disorder and Schizophrenia Working Group of the Psychiatric Genomics Consortium. Electronic address: [email protected]; Bipolar Disorder and Schizophrenia Working Group of the Psychiatric Genomics Consortium. Genomic Dissection of Bipolar Disorder and Schizophrenia, Including 28 Subphenotypes. cell. 2018;173(7):1705–1715.e16. https://doi:10.1016/j.cell.2018.05.046

26. Levine J, Panchalingam K, Rapoport A, Gershon S, McClure RJ, Pettegrew JW. Increased cerebrospinal fluid glutamine levels in depressed patients. Biol. Psychiatry. 20 00;47(7):586–593.

27. Losenkov IS, Boyko AS, Levchuk LA, Simutkin GG, Bokhan NA, Ivanova SA. Serum glutamate in patients with depressive disorders as a potential peripheral marker for predicting the effectiveness of therapy. Neurochemistry. 2018;35(4):359–366.

28. Lee PH, Perlis RH, Jung JY, Byrne EM, Rueckert E, Siburian R, Haddad S, Mayerfeld CE, Heath AC, Pergadia ML, Madden PA, Boomsma DI, Penninx BW, Sklar P, Martin NG, Wray NR, Purcell SM, Smoller JW.