Next Article in Journal
Synthesis and Biological Evaluation of 7-Deoxy-Epothilone Analogues
Next Article in Special Issue
Dicarbonyls and Advanced Glycation End-Products in the Development of Diabetic Complications and Targets for Intervention
Previous Article in Journal
In Vitro Anti-Inflammatory and Cytotoxic Effects of Aqueous Extracts from the Edible Sea Anemones Anemonia sulcata and Actinia equina
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Glyoxalase System and Methylglyoxal-Derived Carbonyl Stress in Sepsis: Glycotoxic Aspects of Sepsis Pathophysiology

1
Department of Anesthesiology, Heidelberg University Hospital, 69120 Heidelberg, Germany
2
Department of Medicine I and Clinical Chemistry, Heidelberg University Hospital, 69120 Heidelberg, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2017, 18(3), 657; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms18030657
Submission received: 11 February 2017 / Revised: 8 March 2017 / Accepted: 14 March 2017 / Published: 17 March 2017
(This article belongs to the Special Issue Glyoxalase System in Health and Disease 2017)

Abstract

:
Sepsis remains one of the leading causes of death in intensive care units. Although sepsis is caused by a viral, fungal or bacterial infection, it is the dysregulated generalized host response that ultimately leads to severe dysfunction of multiple organs and death. The concomitant profound metabolic changes are characterized by hyperglycemia, insulin resistance, and profound transformations of the intracellular energy supply in both peripheral and immune cells. A further hallmark of the early phases of sepsis is a massive formation of reactive oxygen (ROS; e.g., superoxide) as well as nitrogen (RNS; e.g., nitric oxide) species. Reactive carbonyl species (RCS) form a third crucial group of highly reactive metabolites, which until today have been not the focus of interest in sepsis. However, we previously showed in a prospective observational clinical trial that patients suffering from septic shock are characterized by significant methylglyoxal (MG)-derived carbonyl stress, with the glyoxalase system being downregulated in peripheral blood mononuclear cells. In this review, we give a detailed insight into the current state of research regarding the metabolic changes that entail an increased MG-production in septicemia. Thus, we point out the special role of the glyoxalase system in the context of sepsis.

1. Introduction

1.1. Sepsis: Definitions, Incidence, Outcome, and Economic Relevance

Sepsis, defined as a “life-threatening organ dysfunction caused by a dysregulated host response to infection” [1], is characterized by an increasing incidence within the last decades and remains one of the leading causes of death in intensive care units (ICUs) [2,3,4,5,6]. Although mortality rates declined [7] after the implementation of treatment protocols according to the recommendations of the Surviving Sepsis Campaign [8,9,10], overall mortality remains high with rates of 20%–50% causing considerable medical and economic challenges for both health systems and society [2,3,4,5,6].

1.2. Traditional Pathophysiology of Sepsis: PAMPs, PRRs and Reactive Metabolites

Although sepsis is caused by fungal, viral or bacterial infections, the dysregulated generalized host response ultimately leads to severe dysfunction of multiple organs and death [1]. While the underlying pathophysiological mechanisms have not been fully elucidated yet, the innate immune system seems to dominate the early phase of the disease. Pattern recognition receptors (PRR) on immune and endothelial cells recognize unspecific and highly conserved antigens before launching a diversified stress response. These unspecific and highly conserved antigens are called pathogen-associated molecular patterns (PAMPs) and include bacterial cell wall components, such as lipopolysaccharide (LPS) or peptidoglycan (PGN). The resulting cell activation leads to degranulation, phagocytosis, upregulation of adhesion molecules, the release of pro-inflammatory cytokines, such as interleukin 1 (IL-1), IL-6 and tumor necrosis factor α (TNF-α), in addition to the activation of complement pathways causing a pro-coagulatory state [11,12,13,14,15]. Furthermore, studies observed a massive formation of reactive oxygen (ROS; e.g., superoxide) as well as nitrogen (RNS; e.g., nitric oxide) species [11,12,13,14,15]. In addition to ROS and RNS, reactive carbonyl species (RCS) represent another crucial group of highly reactive metabolites [16], which has not been a focus in sepsis research until recently. However, RCS might be a crucial component of the observed changes resulting from the response to an infectious stimulus. In normal physiology, the inflammatory response is strictly limited to the infectious focus with a well-balanced production and elimination of radicals and cytokines [17]. In contrast, during sepsis the body enters a vicious cycle, resulting in a substantial imbalance of these physiological conditions. Altogether these changes have been termed as the systemic inflammatory response syndrome (SIRS) [18]. Although this term was abolished in the latest definition of sepsis [1], it accurately describes the clinical correlations of the early phase of the host response to an infectious stimulus, which is often accompanied by a hyperdynamic cardiovascular circulatory state, fever and changes in the white blood cell count [14,19]. At the same time a counter-regulatory anti-inflammatory response is launched, including an increase in IL-10 release and a remarkable reprogramming of immune cells, partially resulting in immune cell apoptosis and autophagy [20]. Although this compensatory anti-inflammatory response syndrome (CARS) is initiated shortly after the onset of sepsis and its signaling cascades might have an influence on the metabolism in the early phase of sepsis, the first days of the disease are still characterized by the above mentioned pro-inflammatory disease state. The concomitant profound metabolic changes are orchestrated by an interaction of immunoinflammatory and neuroendocrine stress responses [21].

2. Different Aspects of the Stress Response in Sepsis

2.1. Neuroendocrine Reactions in Sepsis

The neuroendocrine response is dominated by activation of the sympathetic nervous system and the hypothalamic-pituitary axis [21]. Among the variety of “stress” hormones that are consequently released, the most important are probably cortisol, epinephrine, norepinephrine, vasopressin in addition to both insulin and glucagon [13,14,15] (Table 1).
Consequently, the first phase of sepsis is often characterized by a hyper-hemodynamic state. Moreover, the excessive release of these hormones result in the metabolic status being substantially changed. The induction of both hepatic gluconeogenesis, glycogenolysis and glycolysis is predominantly triggered by therapeutic and endogenous epinephrine and norepinephrine [22]. Furthermore, norepinephrine increases the supply of glycerol by activating peripheral lipolysis. Cortisol additionally worsens the hyperglycemic state by not only stimulating hepatic gluconeogenesis, but also inducing insulin resistance and thus decreasing glucose uptake by muscles and adipose tissue [22,23]. This induced insulin resistance is further amplified by inflammatory mediators, predominantly TNF-α, IL-1, IL-6 and C-reactive protein [22].

2.2. Metabolic Changes in Sepsis: Hyperglycemia, Insulin Resistance and Respiratory Chain Uncoupling

The immunoinflammatory and neuroendocrine responses to an infectious stimulus lead to a distinct catabolic metabolism co-existing with insulin resistance. As a result, hyperglycemia can frequently be observed in patients with sepsis [22,24]. It is important to note that hyperglycemia, especially severe hyperglycemia, is known to be associated with increased mortality and morbidity in critically ill patients [25,26,27,28,29]. Nevertheless, a large randomized trial targeting intensive glucose control as a means of normalizing blood sugar (81–101 vs. >180 mg/dL) observed an increased mortality in patients with sepsis [30]. Although this increased mortality was suspected to be due to the higher rate of accidental overdosing causing hypoglycemic episodes, it is possible that the excess supply of glucose is a physiological adaption to stress in critical illness. This might be due to the fact that mitochondrial glucose utilization is impaired in many tissues during systemic inflammation [31,32]. One reason might be that the mitochondrial electron transport chain is uncoupled or otherwise altered in these tissues during systemic inflammation [32], with a subsequent large dependence on glycolysis for their energy supply. However, since glycolysis is less effective in producing adenosine triphosphate (ATP) in comparison to oxidative phosphorylation (glycolysis produces only 2 ATP compared with 30 ATP from the citrate circle), large amounts of glucose are temporarily needed in order to cover energy demands. In parallel, a deranged microcirculation further impedes the glucose supply in sepsis. Therefore, the host is forced to provide high blood glucose levels in order to maintain a sufficient glucose gradient, which acts as the driving force for transmembranous glucose transporters (GLUT) [22,23,33]. Both non-insulin-dependent glucose uptake via GLUT1, 2 or 3 and insulin-dependent glucose uptake via GLUT4 rely on concentration gradients, necessitating higher glucose concentrations in the bloodstream than in the cells [34]. While peripheral tissues develop a significant insulin resistance during the course of sepsis due to the downregulation of GLUT4 [35] and the impairment of post-receptor signaling pathways (via the phosphorylation of the insulin receptor, insulin receptor substrate 1 (IRS-1) and MAP kinase) [36], GLUT1 seems to be upregulated in brain cells of mice following thermal injury or infection with Pseudomonas aeruginosa, [37]. Similarly, white blood cells increase the expression of GLUT1, 3 and 4 after being activated [38]. In conclusion, a glucose balance that is adaptive to the necessities of the first phase of the immune response to an infectious stimulus is achieved by high blood sugar levels during systemic inflammation, upregulation of GLUT1 and GLUT3 in addition to the downregulation of GLUT4. These changes facilitate the redirection of glucose away from peripheral tissues towards immune cells and the nervous system [23].
Since hexokinase (the rate limiting enzyme of glycolysis) is saturated under physiological glucose concentrations, glycolytic flux can only be increased via one of the following mechanisms: (1) increase of the hexokinase activity; (2) increase of hexokinase expression; and (3) hexokinase independent glucose-6-phospate generation. All three mechanisms might be of relevance in sepsis, although there might be tissue-specific differences. First, an increase in hexokinase activity was observed in the renal cortical cells of mice following LPS exposure [39]. A similar increase in hexokinase activity was observed in immune cells [40]. Secondly, LPS is known to activate the hypoxia-inducible factor 1α (HIF1α), which in turn induces the upregulation of several glycolytic enzymes, including hexokinase, in immune cells, heart and peripheral tissues [41,42]. Finally, the metabolic shift in sepsis towards a stress state includes high cortisol levels resulting in increased rates of hepatic gluconeogenesis [13], which might be an alternative source of glucose-6-phospate in sepsis.

2.3. Immunometabolism in Sepsis: The “Warburg Effect” and Its Consequences for Immune Cells

The metabolic switches in immune cells during systemic inflammation are of special importance. Aside from changes in utilization of fatty and amino acids, the changes in glucose metabolism seem to be of special importance for the immune response [43]. Thereby, the switch from mitochondrial oxidative phosphorylation to cytosolic glycolysis being the major source of ATP constitutes a hallmark change in all activated immune cells. This phenomenon, known as the “Warburg effect”, was first described in tumor cells [44] and the underlying concept was subsequently extended to include immune cells, such as monocytes, macrophages, dendritic cells [45,46], B cells [47], T cells [48] and natural killer cells [49] during activation [50] (Figure 1).
Although glycolysis is quite inefficient with regard to ATP production in comparison to the oxidative phosphorylation, it provides valuable biosynthetic intermediates and co-enzymes urgently needed for core activities of activated immune cells like cell growth, proliferation or cytokine production [43]. Important examples of processes supported by glycolysis include the reduction of nicotinamide adenine dinucleotide (NAD+) to NADH, which is used by numerous enzymes as a co-factor, and the provision of biosynthetic intermediates for the synthesis of ribose needed for nucleotide production [43]. Through the latter pathway, enhanced glycolysis is intimately linked to an upregulation of the pentose phosphate pathway, providing not only nucleotide intermediates but also NADPH needed for the so-called respiratory burst (discussed below). In a recent review [43], O’Neill et al. gathered molecular insights into the signaling pathways, which are involved in the upregulation of glycolysis during immune cell activation, suggesting important roles for not only hypoxia-inducible factor 1α (HIF1α) [41] and NFκB [45], but also TANK-binding kinase 1 (TBK1), inhibitor of NFκB kinase ε (IKKε) and hexokinase 2 [40].
Aside from its role in the adaption of activated immune cells to fit with particular energy demands of the tissues they move into, glycolysis seems to be involved in further cell line-specific signaling pathways promoting inflammation and immune cell differentiation [43]. For example, it was shown in macrophages that the pyruvate kinase isoenzyme M2 (PKM2), which normally catalyzes the last step of glycolysis and thus regulates its flux, is able to translocate into the nucleus, where it interacts with HIF1α. This results in an upregulation of HIF1α-dependent genes, including IL-1β [41,43,51,52]. Inhibiting this PKM2 translocation into the nucleus results in a switch of the macrophage from the pro-inflammatory M1 phenotype to the rather anti-inflammatory M2 phenotype. In addition, it was shown that the production of IL-1β in macrophages is increased by another mechanism closely connected to glycolysis. Moon et al. demonstrated that the glycolytic enzyme hexokinase is an activator of the nucleotide-binding domain (NOD)-, leucine-rich repeat (LRR)- and pyrin domain-containing 3 (NLRP3) inflammasome, which in turn regulates caspase-1 [53]. Caspase-1 than produces IL-1β, IL-18 and may induce pyroptosis.
Consistent with these findings, another crosslink between glucose metabolism, inflammatory status and phenotype of immune cells was demonstrated in T cells [54,55,56]. In addition, it was shown that the inhibition of glycolysis is able to induce the conversion of T helper 17 (TH17) cells to regulatory T (Treg) cells [57]. Thus, a HIF1α–dependent transcriptional program seemed to determine the glycolytic activity [57]. HIF1α in turn required signaling through mechanistic target of rapamycin (mTOR), making mTOR a key regulator of cellular metabolism [57]. At the same time, mTOR was found to compromise both survival and lineage stability in Treg cells [40,43,58]. Another seminal finding concerning metabolic immune regulation is that the central glycolytic enzyme glyceraldehyde 3-phosphate dehydrogenase (GAPDH) seems to control the translation of interferon-γ (INFγ) in T helper 1 (TH1) cells [59,60]. While GAPDH binds and thus blocks INFγ-encoding mRNA in resting TH1 cells, this blocking effect is weakened in activated TH1 cells. This is due to the fact that an increased glycolytic flux causes a dissociation of GAPDH from the mRNA, resulting in an enhanced translation of the INFγ-encoding mRNA[61].
Combining all of this, hyperglycemia seems to be an adequate stress reaction on one hand, ensuring not only the energy supply of critical organs in systemic inflammation but also enabling the immune system to enter in a pro-inflammatory state and to carry out its designated functions. On the other hand, once a systemic inflammation becomes dysregulated, it might fuel the vicious circle leading to multi-organ dysfunction and death. Thereby, reactive intermediates such as ROS, RNS, and RCS might play a pivotal role.

2.4. Formation of ROS and RNS in Sepsis

As part of the innate immune response to pathogens, phagocytes such as neutrophils or macrophages start to produce massive amounts of superoxide anions (O2) and hydrogen peroxide (H2O2) in order to kill the invaders. Nicotinamide adenine dinucleotide phosphate (NADPH)-oxidases are a major source of these ROS during this process called “respiratory burst” [62]. This respiratory burst results in excessively increased oxygen consumption [63]. Moreover, ROS produced during the respiratory burst seem to act as paracrine agents, both activating adjacent endothelial cells but also altering their function [64,65]. In addition to a large variety of changes, endothelial cells in turn increase intracellular ROS and RNS production. In addition to an NADPH-oxidase-dependent production of ROS, the mitochondrial formation of ROS via the uncoupling of respiratory chain complexes seems to be of great importance [64,65]. Aside from an increased production of nitric oxide (NO) through the inducible NO-synthase (iNOS) [66,67], several other pathways perpetuate ROS production and inflammation. On the one hand, NO stimulates the mitochondrial formation of H2O2 and O2 via the inhibition of the cytochrome-c oxidases [68]. On the other hand, H2O2 induces the upregulation of iNOS via an activation of NFκB [69]. Furthermore, NO reacts with H2O2 to create peroxynitrite and other RNS [70]. Moreover, there is evidence that the mitochondrial complexes I and IV are susceptible to a deterioration by RNS [71,72], which in turn further compromises mitochondrial respiration [73]. Consistent with these findings, different animal models support the finding of altered mitochondrial function in sepsis. Although an impairment of mitochondrial function can be observed in the heart [74,75], liver [76,77] and intestine [78,79], the pathophysiological relevance remains uncertain [32]. Furthermore, Brealy et al. found a reduced mitochondrial complex I activity in the skeletal muscle from biopsies of non-surviving septic patients in comparison to survivors and post-operative controls (at 24 h after ICU admission) [80]. Sjövall et al. observed increased mitochondrial respiration in platelets of septic patients [81,82]. However, oxidative phosphorylation became less effective in these platelets with regard to ATP production, which is due to uncoupling [81]. Correspondingly, the increase in mitochondrial respiration during the early phase of systemic inflammation was inversely correlated with survival [82]. It is important to note that ROS are not only essential microbicides [83] but also represent potent activators of the innate immune system [72,84,85,86,87,88,89,90]. Taking this into consideration, mitochondrial dysfunction in sepsis might be part of the stress response in systemic inflammation. From this point of view, the described shift from the mitochondrial citrate cycle to the cytosolic glycolysis as the predominant source of ATP might be both a sequela of impaired mitochondrial function as well as a necessary measure to permit mitochondrial uncoupling for fueling inflammation in order to fight against the pathogen. Thus, hyperglycemia-induced changes in mitochondrial function in addition to the production of ROS and RNS are facets of a reasonable stress response becoming dysregulated in sepsis. Consistent with this, it was shown that moderate levels of ROS act as second messengers in signaling cascades and gene regulation, including nuclear factor 2 (Nrf2), nuclear factor kappa B (NFκB) and activator protein 1 (AP-1). In comparison, high levels of ROS induce apoptosis or even cause necrosis [91,92,93,94,95,96,97].

2.5. Reactive Carbonyl Species (RCS)—An Overlooked Group of Reactive Metabolites

ROS and RNS are very unstable and their deleterious effects are limited to the location where they are created in addition to the direct surroundings. A third group of more stable but still highly reactive metabolites might help to complete the picture of systemic inflammation. Reactive carbonyl species (RCS) [16] are a heterogeneous group of low molecular carbonyls, which are able to interact with a variety of biomolecules, such as proteins, deoxyribonucleic acid (DNA) of phospholipids, to cause an increased formation of advanced glycation end-products (AGEs). These modification processes involve structural changes, which result in functional changes, dysfunction or a total loss of function of the original protein [98,99]. These damaging effects are comparable to those of ROS, whereas RCS are more stable and thus are able to act systemically [100]. Methylglyoxal (MG) represents a RCS with particular pathophysiological relevance. In mammalian metabolism, MG is formed predominantly as a side product of glycolysis by non-enzymatic degradation of triosephosphates, glyceraldehyde-3-phosphate (GAP) and dihydroxyacetonephosphate (DHAP) [100,101]. Usually, 0.05% to 0.1% of the glycolytic flux is converted into MG [102], resulting in concentrations of 50–150 nM in human plasma [103,104]. Since MG is produced proportional to the glycolytic flux, it is not surprising that elevated MG plasma levels can be found in diseases characterized by altered glycolytic flux (e.g., diabetes mellitus). Moreover, MG seems to be more than a simple surrogate for the unbalanced glucose metabolism in diabetes mellitus, since it is clearly associated with the appearance of diabetic sequelae such as nephropathy and retinopathy [105,106]. Additionally, Bierhaus et al. proved its causal contribution to the development of diabetic neuropathy [107].

2.6. MG-Derived Carbonyl Stress in Sepsis—Diagnostic Value, Prognostic Value and Main Source of Formation

With regards to MG-derived carbonyl stress in critical illness, we were able to show in an earlier study that septic patients feature significantly higher plasma levels of MG compared to postoperative controls and healthy volunteers [108]. Moreover, MG outmatched the established markers of inflammation and infection, such as procalcitonin (PCT), C-reactive protein (CRP), soluble cluster of differentiation (CD) 14 subtype and IL-6, with regards to early and effective detection of sepsis in that study. Furthermore, we identified MG as an independent predictor of mortality in sepsis [108]. Elevated MG plasma levels seem to be inevitable in the chain of reasoning, considering that hyperglycemia is a characteristic feature of sepsis and MG derives from metabolic disease states, which are characterized by an increased glycolytic flux. Consistent with this, some authors interpret human MG-production as an essential defense mechanism of host neutrophils against invading bacteria [109]. However, other studies have shown that bacteria (e.g., Escherichia coli) also produce MG [110,111,112,113] and might therefore be the main source of carbonyl stress in bacterial sepsis. MG production via a specialized MG-synthase enables some E. coli strains to temporarily control their carbon flux [113]. Through this method, these bacteria are able to limit the stress of phosphate intermediates in carbon-rich environments, which allows these cells to grow for a limited time. If the environmental conditions change again to being less carbon-rich, these E. coli strains are able to survive due to this method, otherwise these strains die due to the accumulation of MG. Thus, MG-production embodies a high-risk ecological niche for these bacteria as it provides time for adaptation in unfavorable surrounding conditions [113]. It is still unknown if the main part of the observed MG-kinetics in patients suffering from septic shock are of endogenous or exogenous origin. However, data from our own workgroup strongly support the hypothesis arguing in favor for an endogenous source of carbonyl stress during systemic inflammation. This is due to the fact that, in addition to elevated MG plasma levels in postoperative controls of the above mentioned study [108], patients following liver transplantation [114] and patients suffering from severe trauma [115] are also characterized by significantly increased MG plasma levels. Both these patient groups clearly feature settings of sterile inflammation. Future research projects need to determine whether MG has a causal impact on the course of disease or whether it is just an epiphenomenon following stress hyperglycemia. However, a causal contribution of MG to the detrimental effects of a dysregulated systemic inflammation seems to be reasonable.

2.7. Pathophysiology of MG-Derived Carbonyl Stress in Sepsis

Under physiological conditions, MG reacts with modified arginine, lysine and cysteine residues in proteins [116]. Of these proteins, arginine seems to be the most susceptible. Ahmed et al. demonstrated that MG modifies human serum albumin in vivo [98]. In 92% of the albumin modifications, arginine residues are modified resulting in the formation of hydroimidazolone Nδ-(5-hydro-5-methyl-4-imidazolon-2-yl)-ornithine with a loss of the positive charge of the side chain. Furthermore, they were able to localize a hotspot of modification at Arg-410 via tryptic peptide mapping. Although the clinical relevance of these findings is not clear, it is noteworthy that Arg-410 is located at the drug-binding site II and thus represents the active site of albumin-associated esterase activity [98]. Modified albumin was shown to induce the synthesis and secretion of IL-1β in addition to macrophage colony stimulating factor in human monocytic THP-1 cells [117,118]. In parallel, MG-modified arginine residues act as chemoattractants triggering receptor-mediated endocytosis and degradation [119]. Additionally, Du et al. demonstrated that following MG-mediated oxidation, the protein serine/threonine kinase “Raf-1”, a crucial enzyme for cell growth and development, is recognized as misfolded, which in turn results in ubiquitination and proteasome-dependent proteolysis [120]. Further cell damage can result from the interaction of RCS with DNA, forming the nucleotide AGEs “GdG” and “MGdG” and resulting in strand breaks and point mutations [121]. Meanwhile, it was shown that the plasma concentration of MG and the blood levels of hydroimidazolone (MG-H1) are correlated with the degree of tissue damage and disease progression in diabetes mellitus [105,106,122,123,124]. It was hypothesized that this observed association between high MG/ MG-H1 levels and the progression of diabetic retinopathy and nephropathy [105,106] might be caused by microvascular damage. This theory is based on the formation of MG-H1 at critical functional sites in type IV collagen [125]. Dobler et al. showed that these modifications induce anoikis of endothelial cell and impair angiogenesis [99]. Similar mechanisms are described in mesangial cells [126] and peripheral neurons [127]. Moreover, concerning diabetic neuropathy, Bierhaus et al. [107] showed that the administration of a specifically designed MG scavenger (GERP10, [128]) lowers the plasma levels of MG and MG-H1 and is thereby able to reduce the severity of clinically apparent hyperalgesia.
Regarding critical disease states, the influence of carbonyl stress on mitochondria and its relevance for ROS formation might be of special interest. The most probable theory seems to be that ROS, RNS, and RCS are a self-perpetuating system in which one factor stimulates the other [100,129,130]. Aborno et al. found that MG concentrations were significantly augmented during hydrogen peroxide-induced necrotic cell death [131]. The underlying mechanism involves the inhibition of downstream glycolytic enzymes resulting in an accumulation of glyceraldehyde 3-phosphate (GA3P) and dihydroxyacetone phosphate (DHAP) [132]. Conversely, decreased MG concentrations are associated with decreased ROS levels [131,133]. At the same time, RCS might profoundly impair mitochondrial function. For example, Morcos et al. suggested that mitochondrial proteins feature a particular vulnerability to MG-derived carbonyl stress [128]. In addition, Chan et al. showed that MG treatment increases ROS levels in osteoblasts and induces apoptosis via the mitochondrial pathway [134]. Likewise, Seo et al. treated human hepatic cells in vitro with MG and demonstrated a significantly increased permeability of the mitochondrial membrane under treatment [135]. Consistent with this, they found increased markers for liver cell damage (e.g., alanine transferase (ALAT) and aspartate aminotransferase (ASAT) in mice following treatment with MG. However, the transferability of their results into a clinical setting needs to be analyzed due to high treatment doses in vitro (3 and 10 mM) and in vivo (400 mg/kg bodyweight) [135].
In summary, MG-derived carbonyl stress might significantly alter crucial factors of the immune response to infection. Thus, it might influence not only the effector functions of immune cells but also the integrity of the endothelium and basic metabolic pathways, such as glycolysis and mitochondrial respiration. As a result, it might accelerate the development of sepsis-induced multi-organ dysfunction. However, the quantitative impact and the clinical relevance for the progression of the disease remain unclear and need to be specified.

2.8. Regulation of MG-Derived Carbonyl Stress in Sepsis

Mammals primarily detoxify MG via the glyoxalase pathway, which involves enzymes such as glyoxalase-I (GLO-I) and glyoxalase-II (GLO-II) [136,137]. These enzymes metabolize MG into lactate by using NADPH and glutathione [137,138]. Glyoxalase was first described in 1913 [139,140], but despite a hundred years of research [141], little is known about the actual mechanisms of its regulation. However, this way of MG detoxification seems to be severely disrupted in critical illness (Figure 2).
First of all, systemic inflammation is characterized by massive amounts of ROS, which need to be detoxified. However, a wide range of antioxidant mechanisms rely on glutathione as a co-substrate [142]. The resulting increased glutathione consumption consecutively impairs the glyoxalase system. Secondly, the ROS “hyperoxide” is known to directly inhibit GLO-1 [143]. Consistent with this, we previously found significantly increased levels of both ROS and RCS in critically ill patients compared to post-operative controls and healthy volunteers [108]. Moreover, in patients suffering from septic shock, the expression of GLO-1 was shown to be impaired [108]. These findings could be seminal, since the extent of GLO-1 expression was associated with the outcome of diabetes mellitus (DM) in rodent models [144]. It was postulated that the downregulation of GLO-1 is induced by an activation of the receptor for advanced glycation end-products (RAGE) [145,146,147,148]. Although the underlying pathway remains to be identified, it is well known that RAGE activates NFκB [149]. Moreover, GLO-1 expression has less downregulation in RAGE(−/−) mice with DM than in controls [150]. Moreover, RAGE(−/−) mice are characterized by having significantly improved survival in comparison to wild type controls in a murine model of abdominal sepsis (cecal ligation and puncture) [151]. Conversely, we found an elevated monocytic RAGE expression as well as significantly increased plasma levels of soluble RAGE (sRAGE) in both patients with septic shock [152] and patients following severe trauma [115]. Consistent with an increased MG-derived AGE formation during critical illness, these findings suggest a regulatory circuit involving MG-AGE-RAGE-GLO-1-MG, which results in deleterious formation of ROS, RNS, and RCS in systemic inflammation.
Recently, this hypothesis was challenged in the following way: Van Herreweghe et al. suggested that the regulation of GLO-I activity may be dependent on post-translational modification [153]. They demonstrated that protein kinase A (PKA) phosphorylates the active center of GLO-I and thus alters its function in fibrosarcoma cells. Following this phosphorylation, an accelerated formation of MG-derived AGEs as well as an increased rate of caspase-independent necrotic cell death was observed [153]. However, the informative value of these results is limited, since it is unknown to what extent the results in fibrosarcoma cells are representative for non-cancer cells in inflammation. To begin with, the activity of PKA is regulated by intracellular concentrations of cyclic adenosine monophosphate (cAMP). Cyclic AMP in turn represents the second messengers of many G-protein mediated hormone and cytokine responses [154] (e.g., induction of hepatic glycogenolysis after β-adrenergic stimulation [155]), processes of cell differentiation, cytoskeletal remodeling and proliferation in addition to apoptosis and immune modulation [156]. While AGE formation is thought to be part of pro-inflammatory signaling cascades (via RAGE and NFκB), recent in vivo studies suggest that globally high cAMP concentrations have moderate immunosuppressive effects [157,158]. Furthermore, cAMP-impaired neutrophil chemotaxis in vitro [159] decreased the release of histamine and leukotrienes from basophils or mast cells [160], which significantly reduced ROS formation [161,162]. Since intracellular cAMP measurements in septic patients are hard to perform, little is known about global cAMP levels in sepsis. Increased levels can be assumed in cells harboring adrenergic receptors, whereas cAMP-associated immunosuppressive effects obviously do not prevail, especially in the early phase of sepsis. Thus, it seems questionable whether cAMP levels are globally elevated in sepsis and the observed activation of PKA in fibrosarcoma cells (which in turn downregulates GLO-1 by phosphorylation) [153] can be extrapolated entirely to all cell types of septic patients. It instead appears that cAMP/PKA-dependent GLO regulation in sepsis needs to be evaluated in a more sophisticated manner, since cAMP/PKA signaling is known to be very specific for different tissues, cells and even subcellular units [163,164]. Specific A kinase anchoring proteins (AKAPs) direct PKA signaling to the desired site of action [165]. For NFκB, both stimulating and inhibitory mechanisms of PKA action have been identified [166]. In conclusion, both a RAGE-dependent as well as a PKA-dependent GLO-I regulation might be of value in sepsis. However, a detailed investigation needs to further illuminate the possible interactions and clinical relevance of these pathways. It is important to note that Morgenstern and colleagues recently confirmed the existence of an alternative pathway for the detoxification of MG in GLO-deficient neuronal Schwann cells [167]. In this study, they demonstrated that a knockout of GLO-1 is associated with a compensatory increase of MG detoxification via aldo-keto reductases (AKR), specifically AKR1b3. Whether this mechanism plays a relevant role in systemic inflammation remains to be clarified in further studies.

3. Conclusive Remarks

In summary, sepsis is not only characterized by mass production of cytokines and a dysregulated immune reaction, but also features a significant metabolic stress response. Amongst others, this includes drastic neurohumoral changes, profound alterations in glucose utilization, and a massive release of ROS, RNS, and RCS. Mitochondria might be of major importance in that context, although it is not absolutely clear whether their altered function is the result or the origin of reactive metabolites, such as ROS, RNS, and RCS. In a similar manner, sepsis-associated hyperglycemic metabolic states impair mitochondrial function via the formation of reactive metabolites but are necessary to maintain energy supply in times of impaired mitochondrial function. MG represents a highly reactive RCS and was shown to exhibit a distinct increase in the first hours of sepsis, making MG a more suitable discriminatory marker between patients with sepsis and patients suffering from sterile inflammation in comparison to conventional inflammation markers. Although MG was causally linked to late diabetic sequelae, there is no strong evidence that MG is linked to disease patterns of systemic inflammation and sepsis. However, since MG formation is closely linked to ROS and RNS formation, a causal contribution to the progression of the disease seems to be reasonable. Thus, a modulation of MG-derived carbonyl stress could be a suitable approach for the development of new therapeutic options in sepsis. This could be either realized via a direct scavenging of MG (e.g., GERP-10) or indirectly via a modulation of glyoxalase activity. However, the latter presumes a more detailed knowledge about the (patho)physiology of the glyoxalase-pathway in sepsis, which needs to be addressed in future research projects.

Acknowledgments

We acknowledge the financial support of the Deutsche Forschungsgemeinschaft and Ruprecht-Karls-Universität Heidelberg within the funding programme Open Access Publishing. Thomas Fleming, Jakob Morgenstern and Peter P. Nawroth are supported by SFB1118. We thank Anian Mair for supporting literature research.

Author Contributions

Thomas Schmoch, Florian Uhle, Benedikt H. Siegler, Thomas Fleming, Jakob Morgenstern, Peter P. Nawroth, Markus A. Weigand and Thorsten Brenner contributed to design this review and to write substantial parts of it, according to own expertise. All authors have read and approved the final version of the review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Singer, M.; Deutschman, C.S.; Seymour, C.W.; Shankar-Hari, M.; Annane, D.; Bauer, M.; Bellomo, R.; Bernard, G.R.; Chiche, J.-D.; Coopersmith, C.M.; et al. The Third International Consensus Definitions for Sepsis and Septic Shock (Sepsis-3). J. Am. Med. Assoc. 2016, 315, 801–810. [Google Scholar] [CrossRef] [PubMed]
  2. Martin, G.S.; Mannino, D.M.; Eaton, S.; Moss, M. The epidemiology of sepsis in the United States from 1979 through 2000. N. Engl. J. Med. 2003, 348, 1546–1554. [Google Scholar] [CrossRef] [PubMed]
  3. Bauer, M.; Brunkhorst, F.; Welte, T.; Gerlach, H.; Reinhart, K. Sepsis: Update on pathophysiology, diagnostics and therapy. Anaesthesist 2006, 55, 835–845. [Google Scholar] [CrossRef] [PubMed]
  4. Engel, C.; Brunkhorst, F.M.; Bone, H.-G.; Brunkhorst, R.; Gerlach, H.; Grond, S.; Gruendling, M.; Huhle, G.; Jaschinski, U.; John, S.; et al. Epidemiology of sepsis in Germany: Results from a national prospective multicenter study. Intensive Care Med. 2007, 33, 606–618. [Google Scholar] [CrossRef] [PubMed]
  5. Fleischmann, C.; Thomas-Rueddel, D.; Hartmann, M.; Welte, T.; Heublein, S.; Dennler, U.; Reinhart, K. Hospital Incidence and Mortality Rates of Sepsis: An Analysis of Hospital Episode (DRG) Statistics in Germany from 2007 to 2013. Dtsch. Ärztebl. 2016, 113, 159–166. [Google Scholar]
  6. SepNet Critical Care Trials Group. Incidence of severe sepsis and septic shock in German intensive care units: The prospective, multicentre INSEP study. Intensive Care Med. 2016. [Google Scholar] [CrossRef]
  7. Kaukonen, K.-M.; Bailey, M.; Suzuki, S.; Pilcher, D.; Bellomo, R. Mortality Related to Severe Sepsis and Septic Shock among Critically ill Patients in Australia and New Zealand, 2000–2012. J. Am. Med. Assoc. 2014, 311, 1308–1316. [Google Scholar] [CrossRef] [PubMed]
  8. Dellinger, R.P.; Carlet, J.M.; Masur, H.; Gerlach, H.; Calandra, T.; Cohen, J.; Gea-Banacloche, J.; Keh, D.; Marshall, J.C.; Parker, M.M.; et al. Surviving Sepsis Campaign guidelines for management of severe sepsis and septic shock. Crit. Care Med. 2004, 32, 858–873. [Google Scholar] [CrossRef] [PubMed]
  9. Dellinger, R.P.; Levy, M.M.; Rhodes, A.; Annane, D.; Gerlach, H.; Opal, S.M.; Sevransky, J.E.; Sprung, C.L.; Douglas, I.S.; Jaeschke, R.; et al. Surviving Sepsis Campaign: International guidelines for management of severe sepsis and septic shock, 2012. Intensive Care Med. 2013, 39, 165–228. [Google Scholar] [CrossRef] [PubMed]
  10. SSC Executive Committee. Surviving Sepsis Campaign Bundels—Revised 4/2015; SSC Executive Committee: Wakefield, MA, USA, 2015. [Google Scholar]
  11. Ritthaler, U.; Deng, Y.; Zhang, Y.; Greten, J.; Abel, M.; Sido, B.; Allenberg, J.; Otto, G.; Roth, H.; Bierhaus, A. Expression of receptors for advanced glycation end products in peripheral occlusive vascular disease. Am. J. Pathol. 1995, 146, 688–694. [Google Scholar] [PubMed]
  12. Weigand, M.A.; Hörner, C.; Bardenheuer, H.J.; Bouchon, A. The systemic inflammatory response syndrome. Best Pract. Res. Clin. Anaesthesiol. 2004, 18, 455–475. [Google Scholar] [CrossRef] [PubMed]
  13. Carré, J.E.; Singer, M. Cellular energetic metabolism in sepsis: The need for a systems approach. Biochim. Biophys. Acta 2008, 1777, 763–771. [Google Scholar] [CrossRef] [PubMed]
  14. Hotchkiss, R.S.; Karl, I.E. The pathophysiology and treatment of sepsis. N. Engl. J. Med. 2003, 348, 138–150. [Google Scholar] [CrossRef] [PubMed]
  15. Abraham, E.; Singer, M. Mechanisms of sepsis-induced organ dysfunction. Crit. Care Med. 2007, 35, 2408–2416. [Google Scholar] [CrossRef] [PubMed]
  16. Lide, D.R. CRC Handbook of Chemistry and Physics, 85th ed.; CRC Press: Boca Raton, FL, USA, 2004. [Google Scholar]
  17. Gutteridge, J.M.; Mitchell, J. Redox imbalance in the critically ill. Br. Med. Bull. 1999, 55, 49–75. [Google Scholar] [CrossRef] [PubMed]
  18. Bone, R.C.; Balk, R.A.; Cerra, F.B.; Dellinger, R.P.; Fein, A.M.; Knaus, W.A.; Schein, R.M.; Sibbald, W.J. DEfinitions for sepsis and organ failure and guidelines for the use of innovative therapies in sepsis. The accp/sccm consensus conference committee. American college of chest physicians/society of critical care medicine. Chest 1992, 101, 1644–1655. [Google Scholar] [CrossRef] [PubMed]
  19. Levy, M.M.; Fink, M.P.; Marshall, J.C.; Abraham, E.; Angus, D.; Cook, D.; Cohen, J.; Opal, S.M.; Vincent, J.-L.; Ramsay, G.; et al. International Sepsis Definitions Conference. 2001 SCCM/ESICM/ACCP/ATS/SIS International Sepsis Definitions Conference. Intensive Care Med. 2003, 29, 530–538. [Google Scholar] [CrossRef] [PubMed]
  20. Hotchkiss, R.S.; Monneret, G.; Payen, D. Sepsis-induced immunosuppression: From cellular dysfunctions to immunotherapy. Nat. Rev. Immunol. 2013, 13, 862–874. [Google Scholar] [CrossRef] [PubMed]
  21. Preiser, J.-C.; Ichai, C.; Orban, J.-C.; Groeneveld, A.B.J. Metabolic response to the stress of critical illness. Br. J. Anaesth. 2014, 113, 945–954. [Google Scholar] [CrossRef] [PubMed]
  22. Dungan, K.M.; Braithwaite, S.S.; Preiser, J.-C. Stress hyperglycaemia. Lancet 2009, 373, 1798–1807. [Google Scholar] [CrossRef]
  23. Marik, P.E.; Bellomo, R. Stress hyperglycemia: An essential survival response! Crit. Care 2013, 17, 305. [Google Scholar] [CrossRef] [PubMed]
  24. Marik, P.E.; Raghavan, M. Stress-hyperglycemia, insulin and immunomodulation in sepsis. Intensive Care Med. 2004, 30, 748–756. [Google Scholar] [CrossRef] [PubMed]
  25. Krinsley, J.S. Association between hyperglycemia and increased hospital mortality in a heterogeneous population of critically ill patients. Mayo Clin. Proc. 2003, 78, 1471–1478. [Google Scholar] [CrossRef] [PubMed]
  26. Capes, S.E.; Hunt, D.; Malmberg, K.; Gerstein, H.C. Stress hyperglycaemia and increased risk of death after myocardial infarction in patients with and without diabetes: A systematic overview. Lancet 2000, 355, 773–778. [Google Scholar] [CrossRef]
  27. Gale, S.C.; Sicoutris, C.; Reilly, P.M.; Schwab, C.W.; Gracias, V.H. Poor glycemic control is associated with increased mortality in critically ill trauma patients. Am. Surg. 2007, 73, 454–460. [Google Scholar] [PubMed]
  28. Langley, J.; Adams, G. Insulin-based regimens decrease mortality rates in critically ill patients: A systematic review. Diabetes Metab. Res. Rev. 2007, 23, 184–192. [Google Scholar] [CrossRef] [PubMed]
  29. Wiener, R.S.; Wiener, D.C.; Larson, R.J. Benefits and risks of tight glucose control in critically ill adults: A meta-analysis. J. Am. Med. Assoc. 2008, 300, 933–944. [Google Scholar] [CrossRef] [PubMed]
  30. NICE-SUGAR Study Investigators; Finfer, S.; Chittock, D.R.; Su, S.Y.-S.; Blair, D.; Foster, D.; Dhingra, V.; Bellomo, R.; Cook, D.; Dodek, P.; et al. Intensive versus conventional glucose control in critically ill patients. N. Engl. J. Med. 2009, 360, 1283–1297. [Google Scholar] [PubMed]
  31. Fink, M.P. Cytopathic hypoxia. Mitochondrial dysfunction as mechanism contributing to organ dysfunction in sepsis. Crit. Care Clin. 2001, 17, 219–237. [Google Scholar] [CrossRef]
  32. Fink, M.P. Cytopathic hypoxia and sepsis: Is mitochondrial dysfunction pathophysiologically important or just an epiphenomenon. Pediatr. Crit. Care Med. 2015, 16, 89–91. [Google Scholar] [CrossRef] [PubMed]
  33. Losser, M.-R.; Damoisel, C.; Payen, D. Bench-to-bedside review: Glucose and stress conditions in the intensive care unit. Crit. Care 2010, 14, 231. [Google Scholar] [CrossRef] [PubMed]
  34. Heinrich, P.C.; Müller, M.; Graeve, L. Löffler/Petrides Biochemie und Pathobiochemie; Vollständig Überarbeitete Auflage; Springer: Berlin/Heidelberg, Germany, 2014; Volume 9. [Google Scholar]
  35. Lang, C.H.; Dobrescu, C.; Mészáros, K. Insulin-mediated glucose uptake by individual tissues during sepsis. Metabolism 1990, 39, 1096–1107. [Google Scholar] [CrossRef]
  36. Fan, J.; Li, Y.H.; Wojnar, M.M.; Lang, C.H. Endotoxin-induced alterations in insulin-stimulated phosphorylation of insulin receptor, IRS-1, and MAP kinase in skeletal muscle. Shock Augusta Ga 1996, 6, 164–170. [Google Scholar] [CrossRef]
  37. Gamelli, R.L.; Liu, H.; He, L.K.; Hofmann, C.A. Alterations of glucose transporter mRNA and protein levels in brain following thermal injury and sepsis in mice. Shock Augusta Ga 1994, 1, 395–400. [Google Scholar] [CrossRef]
  38. Maratou, E.; Dimitriadis, G.; Kollias, A.; Boutati, E.; Lambadiari, V.; Mitrou, P.; Raptis, S.A. Glucose transporter expression on the plasma membrane of resting and activated white blood cells. Eur. J. Clin. Investig. 2007, 37, 282–290. [Google Scholar] [CrossRef] [PubMed]
  39. Smith, J.A.; Stallons, L.J.; Schnellmann, R.G. Renal cortical hexokinase and pentose phosphate pathway activation through the EGFR/Akt signaling pathway in endotoxin-induced acute kidney injury. Am. J. Physiol. Ren. Physiol. 2014, 307, F435–F444. [Google Scholar] [CrossRef] [PubMed]
  40. Huynh, A.; DuPage, M.; Priyadharshini, B.; Sage, P.T.; Quiros, J.; Borges, C.M.; Townamchai, N.; Gerriets, V.A.; Rathmell, J.C.; Sharpe, A.H.; et al. Control of PI(3) kinase in Treg cells maintains homeostasis and lineage stability. Nat. Immunol. 2015, 16, 188–196. [Google Scholar] [CrossRef] [PubMed]
  41. Tannahill, G.M.; Curtis, A.M.; Adamik, J.; Palsson-McDermott, E.M.; McGettrick, A.F.; Goel, G.; Frezza, C.; Bernard, N.J.; Kelly, B.; Foley, N.H.; et al. Succinate is an inflammatory signal that induces IL-1β through HIF-1α. Nature 2013, 496, 238–242. [Google Scholar] [CrossRef] [PubMed]
  42. Roberts, D.J.; Miyamoto, S. Hexokinase II integrates energy metabolism and cellular protection: Akting on mitochondria and TORCing to autophagy. Cell Death Differ. 2015, 22, 248–257. [Google Scholar] [CrossRef] [PubMed]
  43. O’Neill, L.A.J.; Kishton, R.J.; Rathmell, J. A guide to immunometabolism for immunologists. Nat. Rev. Immunol. 2016, 16, 553–565. [Google Scholar] [CrossRef] [PubMed]
  44. Hsu, P.P.; Sabatini, D.M. Cancer Cell Metabolism: Warburg and Beyond. Cell 2008, 134, 703–707. [Google Scholar] [CrossRef] [PubMed]
  45. Rodríguez-Prados, J.-C.; Través, P.G.; Cuenca, J.; Rico, D.; Aragonés, J.; Martín-Sanz, P.; Cascante, M.; Boscá, L. Substrate Fate in Activated Macrophages: A Comparison between Innate, Classic, and Alternative Activation. J. Immunol. 2010, 185, 605–614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Krawczyk, C.M.; Holowka, T.; Sun, J.; Blagih, J.; Amiel, E.; DeBerardinis, R.J.; Cross, J.R.; Jung, E.; Thompson, C.B.; Jones, R.G.; et al. Toll-like receptor-induced changes in glycolytic metabolism regulate dendritic cell activation. Blood 2010, 115, 4742–4749. [Google Scholar] [CrossRef] [PubMed]
  47. Doughty, C.A.; Bleiman, B.F.; Wagner, D.J.; Dufort, F.J.; Mataraza, J.M.; Roberts, M.F.; Chiles, T.C. Antigen receptor-mediated changes in glucose metabolism in B lymphocytes: Role of phosphatidylinositol 3-kinase signaling in the glycolytic control of growth. Blood 2006, 107, 4458–4465. [Google Scholar] [CrossRef] [PubMed]
  48. Michalek, R.D.; Gerriets, V.A.; Jacobs, S.R.; Macintyre, A.N.; MacIver, N.J.; Mason, E.F.; Sullivan, S.A.; Nichols, A.G.; Rathmell, J.C. Cutting edge: Distinct glycolytic and lipid oxidative metabolic programs are essential for effector and regulatory CD4+ T cell subsets. J. Immunol. Baltim. Md 1950 2011, 186, 3299–3303. [Google Scholar] [CrossRef] [PubMed]
  49. Donnelly, R.P.; Loftus, R.M.; Keating, S.E.; Liou, K.T.; Biron, C.A.; Gardiner, C.M.; Finlay, D.K. mTORC1-dependent metabolic reprogramming is a prerequisite for NK cell effector function. J. Immunol. Baltim. Md 1950 2014, 193, 4477–4484. [Google Scholar] [CrossRef] [PubMed]
  50. Palsson-McDermott, E.M.; O’Neill, L.A.J. The Warburg effect then and now: From cancer to inflammatory diseases. BioEssays News Rev. Mol. Cell. Dev. Biol. 2013, 35, 965–973. [Google Scholar] [CrossRef] [PubMed]
  51. Luo, W.; Hu, H.; Chang, R.; Zhong, J.; Knabel, M.; O’Meally, R.; Cole, R.N.; Pandey, A.; Semenza, G.L. Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell 2011, 145, 732–744. [Google Scholar] [CrossRef] [PubMed]
  52. Palsson-McDermott, E.M.; Curtis, A.M.; Goel, G.; Lauterbach, M.A.R.; Sheedy, F.J.; Gleeson, L.E.; van den Bosch, M.W.M.; Quinn, S.R.; Domingo-Fernandez, R.; Johnston, D.G.W.; et al. 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. 2015, 21, 65–80. [Google Scholar] [CrossRef] [PubMed]
  53. Moon, J.-S.; Hisata, S.; Park, M.-A.; DeNicola, G.M.; Ryter, S.W.; Nakahira, K.; Choi, A.M.K. mTORC1-Induced HK1-Dependent Glycolysis Regulates NLRP3 Inflammasome Activation. Cell Rep. 2015, 12, 102–115. [Google Scholar] [CrossRef] [PubMed]
  54. Jha, A.K.; Huang, S.C.-C.; Sergushichev, A.; Lampropoulou, V.; Ivanova, Y.; Loginicheva, E.; Chmielewski, K.; Stewart, K.M.; Ashall, J.; Everts, B.; et al. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization. Immunity 2015, 42, 419–430. [Google Scholar] [CrossRef] [PubMed]
  55. Gerriets, V.A.; Kishton, R.J.; Nichols, A.G.; Macintyre, A.N.; Inoue, M.; Ilkayeva, O.; Winter, P.S.; Liu, X.; Priyadharshini, B.; Slawinska, M.E.; et al. Metabolic programming and PDHK1 control CD4+ T cell subsets and inflammation. J. Clin. Investig. 2015, 125, 194–207. [Google Scholar] [CrossRef] [PubMed]
  56. Beier, U.H.; Angelin, A.; Akimova, T.; Wang, L.; Liu, Y.; Xiao, H.; Koike, M.A.; Hancock, S.A.; Bhatti, T.R.; Han, R.; et al. Essential role of mitochondrial energy metabolism in Foxp3+ T-regulatory cell function and allograft survival. FASEB J. 2015, 29, 2315–2326. [Google Scholar] [CrossRef] [PubMed]
  57. Shi, L.Z.; Wang, R.; Huang, G.; Vogel, P.; Neale, G.; Green, D.R.; Chi, H. HIF1α–dependent glycolytic pathway orchestrates a metabolic checkpoint for the differentiation of TH17 and Treg cells. J. Exp. Med. 2011, 208, 1367–1376. [Google Scholar] [CrossRef] [PubMed]
  58. Shrestha, S.; Yang, K.; Guy, C.; Vogel, P.; Neale, G.; Chi, H. Treg cells require the phosphatase PTEN to restrain TH1 and TFH cell responses. Nat. Immunol. 2015, 16, 178–187. [Google Scholar] [CrossRef] [PubMed]
  59. Chang, C.-H.; Curtis, J.D.; Maggi, L.B.; Faubert, B.; Villarino, A.V.; O’Sullivan, D.; Huang, S.C.-C.; van der Windt, G.J.W.; Blagih, J.; Qiu, J.; et al. Posttranscriptional control of T cell effector function by aerobic glycolysis. Cell 2013, 153, 1239–1251. [Google Scholar] [CrossRef] [PubMed]
  60. Mukhopadhyay, R.; Jia, J.; Arif, A.; Ray, P.S.; Fox, P.L. The GAIT system: A gatekeeper of inflammatory gene expression. Trends Biochem. Sci. 2009, 34, 324–331. [Google Scholar] [CrossRef] [PubMed]
  61. Schroder, K.; Hertzog, P.J.; Ravasi, T.; Hume, D.A. Interferon-gamma: An overview of signals, mechanisms and functions. J. Leukoc. Biol. 2004, 75, 163–189. [Google Scholar] [CrossRef] [PubMed]
  62. Dahlgren, C.; Karlsson, A. Respiratory burst in human neutrophils. J. Immunol. Methods 1999, 232, 3–14. [Google Scholar] [CrossRef]
  63. Baldridge, C.W.; Gerard, R.W. The Extra Respiration of Phagocytosis. Am. J. Physiol. Leg. Content 1932, 103, 235–236. [Google Scholar]
  64. Cepinskas, G.; Wilson, J.X. Inflammatory Response in Microvascular Endothelium in Sepsis: Role of Oxidants. J. Clin. Biochem. Nutr. 2008, 42, 175–184. [Google Scholar] [CrossRef] [PubMed]
  65. Ince, C.; Mayeux, P.R.; Nguyen, T.; Gomez, H.; Kellum, J.A.; Ospina-Tascón, G.A.; Hernandez, G.; Murray, P.; de Backer, D.; ADQI XIV Workgroup. THE ENDOTHELIUM IN SEPSIS. Shock Augusta Ga 2016, 45, 259–270. [Google Scholar] [CrossRef] [PubMed]
  66. Santos, S.S.; Brunialti, M.K.C.; Rigato, O.; Machado, F.R.; Silva, E.; Salomao, R. Generation of nitric oxide and reactive oxygen species by neutrophils and monocytes from septic patients and association with outcomes. Shock Augusta Ga 2012, 38, 18–23. [Google Scholar] [CrossRef] [PubMed]
  67. Cobb, J.P. Nitric Oxide Synthase Inhibition as Therapy for Sepsis: A Decade of Promise. Surg. Infect. 2001, 2, 93–101. [Google Scholar] [CrossRef] [PubMed]
  68. Poderoso, J.J.; Carreras, M.C.; Lisdero, C.; Riobó, N.; Schöpfer, F.; Boveris, A. Nitric Oxide Inhibits Electron Transfer and Increases Superoxide Radical Production in Rat Heart Mitochondria and Submitochondrial Particles. Arch. Biochem. Biophys. 1996, 328, 85–92. [Google Scholar] [CrossRef] [PubMed]
  69. Han, Y.-J.; Kwon, Y.-G.; Chung, H.-T.; Lee, S.-K.; Simmons, R.L.; Billiar, T.R.; Kim, Y.-M. Antioxidant Enzymes Suppress Nitric Oxide Production through the Inhibition of NF-κB Activation: Role of H2O2 and Nitric Oxide in Inducible Nitric Oxide Synthase Expression in Macrophages. Nitric Oxide 2001, 5, 504–513. [Google Scholar] [CrossRef] [PubMed]
  70. Cuzzocrea, S.; Mazzon, E.; Paola, R.D.; Esposito, E.; Macarthur, H.; Matuschak, G.M.; Salvemini, D. A Role for Nitric Oxide-Mediated Peroxynitrite Formation in a Model of Endotoxin-Induced Shock. J. Pharmacol. Exp. Ther. 2006, 319, 73–81. [Google Scholar] [CrossRef] [PubMed]
  71. Beltrán, B.; Orsi, A.; Clementi, E.; Moncada, S. Oxidative stress and S-nitrosylation of proteins in cells. Br. J. Pharmacol. 2000, 129, 953–960. [Google Scholar] [CrossRef] [PubMed]
  72. Arulkumaran, N.; Deutschman, C.S.; Pinsky, M.R.; Zuckerbraun, B.; Schumacker, P.T.; Gomez, H.; Gomez, A.; Murray, P.; Kellum, J.A. Mitochondrial Function in Sepsis. Shock 2016, 45, 271–281. [Google Scholar] [CrossRef] [PubMed]
  73. Boulos, M.; Astiz, M.E.; Barua, R.S.; Osman, M. Impaired mitochondrial function induced by serum from septic shock patients is attenuated by inhibition of nitric oxide synthase and poly(ADP-ribose) synthase. Crit. Care Med. 2003, 31, 353–358. [Google Scholar] [CrossRef] [PubMed]
  74. Vanasco, V.; Saez, T.; Magnani, N.D.; Pereyra, L.; Marchini, T.; Corach, A.; Inés Vaccaro, M.; Corach, D.; Evelson, P.; Alvarez, S. Cardiac mitochondrial biogenesis in endotoxemia is not accompanied by mitochondrial function recovery. Free Radic. Biol. Med. 2014, 77, 1–9. [Google Scholar] [CrossRef] [PubMed]
  75. Levy, R.J.; Vijayasarathy, C.; Raj, N.R.; Avadhani, N.G.; Deutschman, C.S. Competitive and noncompetitive inhibition of myocardial cytochrome C oxidase in sepsis. Shock Augusta Ga 2004, 21, 110–114. [Google Scholar] [CrossRef] [PubMed]
  76. Lowes, D.A.; Webster, N.R.; Murphy, M.P.; Galley, H.F. Antioxidants that protect mitochondria reduce interleukin-6 and oxidative stress, improve mitochondrial function, and reduce biochemical markers of organ dysfunction in a rat model of acute sepsis. Br. J. Anaesth. 2013, 110, 472–480. [Google Scholar] [CrossRef] [PubMed]
  77. Brealey, D.; Karyampudi, S.; Jacques, T.S.; Novelli, M.; Stidwill, R.; Taylor, V.; Smolenski, R.T.; Singer, M. Mitochondrial dysfunction in a long-term rodent model of sepsis and organ failure. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2004, 286, R491–R497. [Google Scholar] [CrossRef] [PubMed]
  78. Crouser, E.D.; Julian, M.W.; Dorinsky, P.M. Ileal VO2-O2 alterations induced by endotoxin correlate with severity of mitochondrial injury. Am. J. Respir. Crit. Care Med. 1999, 160, 1347–1353. [Google Scholar] [CrossRef] [PubMed]
  79. King, C.J.; Tytgat, S.; Delude, R.L.; Fink, M.P. Ileal mucosal oxygen consumption is decreased in endotoxemic rats but is restored toward normal by treatment with aminoguanidine. Crit. Care Med. 1999, 27, 2518–2524. [Google Scholar] [CrossRef] [PubMed]
  80. Brealey, D.; Brand, M.; Hargreaves, I.; Heales, S.; Land, J.; Smolenski, R.; Davies, N.A.; Cooper, C.E.; Singer, M. Association between mitochondrial dysfunction and severity and outcome of septic shock. Lancet 2002, 360, 219–223. [Google Scholar] [CrossRef]
  81. Sjövall, F.; Morota, S.; Frostner, E.Å.; Hansson, M.J.; Elmér, E. Cytokine and nitric oxide levels in patients with sepsis—Temporal evolvement and relation to platelet mitochondrial respiratory function. PLoS ONE 2014, 9, e103756. [Google Scholar] [CrossRef] [PubMed]
  82. Sjövall, F.; Morota, S.; Hansson, M.J.; Friberg, H.; Gnaiger, E.; Elmér, E. Temporal increase of platelet mitochondrial respiration is negatively associated with clinical outcome in patients with sepsis. Crit. Care Lond. Engl. 2010, 14, R214. [Google Scholar] [CrossRef] [PubMed]
  83. Babior, B.M. Phagocytes and oxidative stress. Am. J. Med. 2000, 109, 33–44. [Google Scholar] [CrossRef]
  84. Zhou, R.; Yazdi, A.S.; Menu, P.; Tschopp, J. A role for mitochondria in NLRP3 inflammasome activation. Nature 2011, 469, 221–225. [Google Scholar] [CrossRef] [PubMed]
  85. Mao, K.; Chen, S.; Chen, M.; Ma, Y.; Wang, Y.; Huang, B.; He, Z.; Zeng, Y.; Hu, Y.; Sun, S.; et al. Nitric oxide suppresses NLRP3 inflammasome activation and protects against LPS-induced septic shock. Cell Res. 2013, 23, 201–212. [Google Scholar] [CrossRef] [PubMed]
  86. Emre, Y.; Hurtaud, C.; Nübel, T.; Criscuolo, F.; Ricquier, D.; Cassard-Doulcier, A.-M. Mitochondria contribute to LPS-induced MAPK activation via uncoupling protein UCP2 in macrophages. Biochem. J. 2007, 402, 271–278. [Google Scholar] [CrossRef] [PubMed]
  87. Rousset, S.; Emre, Y.; Join-Lambert, O.; Hurtaud, C.; Ricquier, D.; Cassard-Doulcier, A.-M. The uncoupling protein 2 modulates the cytokine balance in innate immunity. Cytokine 2006, 35, 135–142. [Google Scholar] [CrossRef] [PubMed]
  88. West, A.P.; Shadel, G.S.; Ghosh, S. Mitochondria in innate immune responses. Nat. Rev. Immunol. 2011, 11, 389–402. [Google Scholar] [CrossRef] [PubMed]
  89. West, A.P.; Brodsky, I.E.; Rahner, C.; Woo, D.K.; Erdjument-Bromage, H.; Tempst, P.; Walsh, M.C.; Choi, Y.; Shadel, G.S.; Ghosh, S. TLR signaling augments macrophage bactericidal activity through mitochondrial ROS. Nature 2011, 472, 476–480. [Google Scholar] [CrossRef] [PubMed]
  90. Chandel, N.S.; Schumacker, P.T.; Arch, R.H. Reactive Oxygen Species Are Downstream Products of TRAF-mediated Signal Transduction. J. Biol. Chem. 2001, 276, 42728–42736. [Google Scholar] [CrossRef] [PubMed]
  91. Hensley, K.; Robinson, K.A.; Gabbita, S.P.; Salsman, S.; Floyd, R.A. Reactive oxygen species, cell signaling, and cell injury. Free Radic. Biol. Med. 2000, 28, 1456–1462. [Google Scholar] [CrossRef]
  92. Lander, H.M. An essential role for free radicals and derived species in signal transduction. FASEB J. 1997, 11, 118–124. [Google Scholar] [PubMed]
  93. Genestra, M. Oxyl radicals, redox-sensitive signalling cascades and antioxidants. Cell. Signal. 2007, 19, 1807–1819. [Google Scholar] [CrossRef] [PubMed]
  94. Tonks, N.K. Redox redux: Revisiting PTPs and the control of cell signaling. Cell 2005, 121, 667–670. [Google Scholar] [CrossRef] [PubMed]
  95. Halliwell, B.; Gutteridge, J.M.C. Free Radicals in Biology and Medicine; Oxford University Press: Oxford, UK, 2015. [Google Scholar]
  96. Aslan, M.; Ozben, T. Oxidants in receptor tyrosine kinase signal transduction pathways. Antioxid. Redox Signal. 2003, 5, 781–788. [Google Scholar] [CrossRef] [PubMed]
  97. Schreck, R.; Rieber, P.; Baeuerle, P.A. Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-κ B transcription factor and HIV-1. EMBO J. 1991, 10, 2247–2258. [Google Scholar] [PubMed]
  98. Ahmed, N.; Dobler, D.; Dean, M.; Thornalley, P.J. Peptide mapping identifies hotspot site of modification in human serum albumin by methylglyoxal involved in ligand binding and esterase activity. J. Biol. Chem. 2005, 280, 5724–5732. [Google Scholar] [CrossRef] [PubMed]
  99. Dobler, D.; Ahmed, N.; Song, L.; Eboigbodin, K.E.; Thornalley, P.J. Increased dicarbonyl metabolism in endothelial cells in hyperglycemia induces anoikis and impairs angiogenesis by RGD and GFOGER motif modification. Diabetes 2006, 55, 1961–1969. [Google Scholar] [CrossRef] [PubMed]
  100. Kalapos, M.P. The tandem of free radicals and methylglyoxal. Chem. Biol. Interact. 2008, 171, 251–271. [Google Scholar] [CrossRef] [PubMed]
  101. Thornalley, P.J.; Langborg, A.; Minhas, H.S. Formation of glyoxal, methylglyoxal and 3-deoxyglucosone in the glycation of proteins by glucose. Biochem. J. 1999, 344 Pt 1, 109–116. [Google Scholar] [CrossRef] [PubMed]
  102. Veech, R.L.; Raijman, L.; Dalziel, K.; Krebs, H.A. Disequilibrium in the triose phosphate isomerase system in rat liver. Biochem. J. 1969, 115, 837–842. [Google Scholar] [CrossRef] [PubMed]
  103. Rabbani, N.; Thornalley, P.J. Measurement of methylglyoxal by stable isotopic dilution analysis LC-MS/MS with corroborative prediction in physiological samples. Nat. Protoc. 2014, 9, 1969–1979. [Google Scholar] [CrossRef] [PubMed]
  104. Rabbani, N.; Xue, M.; Thornalley, P.J. Dicarbonyls and glyoxalase in disease mechanisms and clinical therapeutics. Glycoconj. J. 2016, 33, 513–525. [Google Scholar] [CrossRef] [PubMed]
  105. Vander Jagt, D.L. Methylglyoxal, diabetes mellitus and diabetic complications. Drug Metabol. Drug Interact. 2008, 23, 93–124. [Google Scholar] [PubMed]
  106. Beisswenger, P.J.; Howell, S.K.; Nelson, R.G.; Mauer, M.; Szwergold, B.S. α-Oxoaldehyde metabolism and diabetic complications. Biochem. Soc. Trans. 2003, 31, 1358–1363. [Google Scholar] [CrossRef] [PubMed]
  107. Bierhaus, A.; Fleming, T.; Stoyanov, S.; Leffler, A.; Babes, A.; Neacsu, C.; Sauer, S.K.; Eberhardt, M.; Schnölzer, M.; Lasitschka, F.; et al. Methylglyoxal modification of Nav1.8 facilitates nociceptive neuron firing and causes hyperalgesia in diabetic neuropathy. Nat. Med. 2012, 18, 926–933. [Google Scholar] [CrossRef] [PubMed]
  108. Brenner, T.; Fleming, T.; Uhle, F.; Silaff, S.; Schmitt, F.; Salgado, E.; Ulrich, A.; Zimmermann, S.; Bruckner, T.; Martin, E.; et al. Methylglyoxal as a new biomarker in patients with septic shock: An observational clinical study. Crit. Care Lond. Engl. 2014, 18, 683. [Google Scholar] [CrossRef] [PubMed]
  109. Zhang, M.M.; Ong, C.Y.; Walker, M.J.; McEwan, A.G. Defence against methylglyoxal in Group A Streptococcus: A role for Glyoxylase I in bacterial virulence and survival in neutrophils? Pathog. Dis. 2016, 74, ftv122. [Google Scholar] [CrossRef] [PubMed]
  110. Baskaran, S.; Rajan, D.P.; Balasubramanian, K.A. Formation of methylglyoxal by bacteria isolated from human faeces. J. Med. Microbiol. 1989, 28, 211–215. [Google Scholar] [CrossRef] [PubMed]
  111. Booth, I.R.; Ferguson, G.P.; Miller, S.; Li, C.; Gunasekera, B.; Kinghorn, S. Bacterial production of methylglyoxal: A survival strategy or death by misadventure? Biochem. Soc. Trans. 2003, 31, 1406–1408. [Google Scholar] [CrossRef] [PubMed]
  112. Cooper, R.A. Metabolism of methylglyoxal in microorganisms. Annu. Rev. Microbiol. 1984, 38, 49–68. [Google Scholar] [CrossRef] [PubMed]
  113. Ferguson, G.P.; Tötemeyer, S.; MacLean, M.J.; Booth, I.R. Methylglyoxal production in bacteria: Suicide or survival? Arch. Microbiol. 1998, 170, 209–218. [Google Scholar] [CrossRef] [PubMed]
  114. Brenner, T.; Fleming, T.H.; Spranz, D.; Schemmer, P.; Bruckner, T.; Uhle, F.; Martin, E.O.; Weigand, M.A.; Hofer, S. Reactive Metabolites and AGE-RAGE-Mediated Inflammation in Patients following Liver Transplantation. Mediat. Inflamm. 2013, 2013, 501430. [Google Scholar] [CrossRef] [PubMed]
  115. Uhle, F.; Lichtenstern, C.; Brenner, T.; Fleming, T.; Koch, C.; Hecker, A.; Heiss, C.; Nawroth, P.P.; Hofer, S.; Weigand, M.A.; et al. Role of the RAGE Axis during the Immune Response after Severe Trauma: A Prospective Pilot Study. Mediat. Inflamm. 2015, 2015, e691491. [Google Scholar] [CrossRef] [PubMed]
  116. Lo, T.W.; Westwood, M.E.; McLellan, A.C.; Selwood, T.; Thornalley, P.J. Binding and modification of proteins by methylglyoxal under physiological conditions. A kinetic and mechanistic study with N α-acetylarginine, N α-acetylcysteine, and N α-acetyllysine, and bovine serum albumin. J. Biol. Chem. 1994, 269, 32299–32305. [Google Scholar] [PubMed]
  117. Westwood, M.E.; Thornalley, P.J. Induction of synthesis and secretion of interleukin 1 β in the human monocytic THP-1 cells by human serum albumins modified with methylglyoxal and advanced glycation endproducts. Immunol. Lett. 1996, 50, 17–21. [Google Scholar] [CrossRef]
  118. Abordo, E.A.; Westwood, M.E.; Thornalley, P.J. Synthesis and secretion of macrophage colony stimulating factor by mature human monocytes and human monocytic THP-1 cells induced by human serum albumin derivatives modified with methylglyoxal and glucose-derived advanced glycation endproducts. Immunol. Lett. 1996, 53, 7–13. [Google Scholar] [CrossRef]
  119. Westwood, M.E.; Argirov, O.K.; Abordo, E.A.; Thornalley, P.J. Methylglyoxal-modified arginine residues—A signal for receptor-mediated endocytosis and degradation of proteins by monocytic THP-1 cells. Biochim. Biophys. Acta 1997, 1356, 84–94. [Google Scholar] [CrossRef]
  120. Du, J.; Zeng, J.; Ou, X.; Ren, X.; Cai, S. Methylglyoxal downregulates Raf-1 protein through a ubiquitination-mediated mechanism. Int. J. Biochem. Cell Biol. 2006, 38, 1084–1091. [Google Scholar] [CrossRef] [PubMed]
  121. Hoon, S.; Gebbia, M.; Costanzo, M.; Davis, R.W.; Giaever, G.; Nislow, C. A global perspective of the genetic basis for carbonyl stress resistance. G3 Bethesda Md 2011, 1, 219–231. [Google Scholar] [CrossRef] [PubMed]
  122. Berner, A.K.; Brouwers, O.; Pringle, R.; Klaassen, I.; Colhoun, L.; McVicar, C.; Brockbank, S.; Curry, J.W.; Miyata, T.; Brownlee, M.; et al. Protection against methylglyoxal-derived AGEs by regulation of glyoxalase 1 prevents retinal neuroglial and vasodegenerative pathology. Diabetologia 2012, 55, 845–854. [Google Scholar] [CrossRef] [PubMed]
  123. Fleming, T.H.; Humpert, P.M.; Nawroth, P.P.; Bierhaus, A. Reactive metabolites and AGE/RAGE-mediated cellular dysfunction affect the aging process: A mini-review. Gerontology 2011, 57, 435–443. [Google Scholar] [CrossRef] [PubMed]
  124. Dhar, A.; Dhar, I.; Desai, K.M.; Wu, L. Methylglyoxal scavengers attenuate endothelial dysfunction induced by methylglyoxal and high concentrations of glucose. Br. J. Pharmacol. 2010, 161, 1843–1856. [Google Scholar] [CrossRef] [PubMed]
  125. Rabbani, N.; Thornalley, P.J. Dicarbonyl stress in cell and tissue dysfunction contributing to ageing and disease. Biochem. Biophys. Res. Commun. 2015, 458, 221–226. [Google Scholar] [CrossRef] [PubMed]
  126. Pedchenko, V.K.; Chetyrkin, S.V.; Chuang, P.; Ham, A.-J.L.; Saleem, M.A.; Mathieson, P.W.; Hudson, B.G.; Voziyan, P.A. Mechanism of perturbation of integrin-mediated cell-matrix interactions by reactive carbonyl compounds and its implication for pathogenesis of diabetic nephropathy. Diabetes 2005, 54, 2952–2960. [Google Scholar] [CrossRef] [PubMed]
  127. Duran-Jimenez, B.; Dobler, D.; Moffatt, S.; Rabbani, N.; Streuli, C.H.; Thornalley, P.J.; Tomlinson, D.R.; Gardiner, N.J. Advanced glycation end products in extracellular matrix proteins contribute to the failure of sensory nerve regeneration in diabetes. Diabetes 2009, 58, 2893–2903. [Google Scholar] [CrossRef] [PubMed]
  128. Morcos, M.; Du, X.; Pfisterer, F.; Hutter, H.; Sayed, A.A.R.; Thornalley, P.; Ahmed, N.; Baynes, J.; Thorpe, S.; Kukudov, G.; et al. Glyoxalase-1 prevents mitochondrial protein modification and enhances lifespan in Caenorhabditis elegans. Aging Cell 2008, 7, 260–269. [Google Scholar] [CrossRef] [PubMed]
  129. Kalapos, M.P.; Littauer, A.; de Groot, H. Has reactive oxygen a role in methylglyoxal toxicity? A study on cultured rat hepatocytes. Arch. Toxicol. 1993, 67, 369–372. [Google Scholar] [CrossRef] [PubMed]
  130. Wu, L. The pro-oxidant role of methylglyoxal in mesenteric artery smooth muscle cells. Can. J. Physiol. Pharmacol. 2005, 83, 63–68. [Google Scholar] [CrossRef] [PubMed]
  131. Abordo, E.A.; Minhas, H.S.; Thornalley, P.J. Accumulation of α-oxoaldehydes during oxidative stress: A role in cytotoxicity. Biochem. Pharmacol. 1999, 58, 641–648. [Google Scholar] [CrossRef]
  132. Hyslop, P.A.; Hinshaw, D.B.; Halsey, W.A.; Schraufstätter, I.U.; Sauerheber, R.D.; Spragg, R.G.; Jackson, J.H.; Cochrane, C.G. Mechanisms of oxidant-mediated cell injury. The glycolytic and mitochondrial pathways of ADP phosphorylation are major intracellular targets inactivated by hydrogen peroxide. J. Biol. Chem. 1988, 263, 1665–1675. [Google Scholar] [PubMed]
  133. Yao, D.; Brownlee, M. Hyperglycemia-induced reactive oxygen species increase expression of the receptor for advanced glycation end products (RAGE) and RAGE ligands. Diabetes 2010, 59, 249–255. [Google Scholar] [CrossRef] [PubMed]
  134. Chan, W.-H.; Wu, H.-J.; Shiao, N.-H. Apoptotic signaling in methylglyoxal-treated human osteoblasts involves oxidative stress, c-Jun N-terminal kinase, caspase-3, and p21-activated kinase 2. J. Cell. Biochem. 2007, 100, 1056–1069. [Google Scholar] [CrossRef] [PubMed]
  135. Seo, K.; Ki, S.H.; Shin, S.M. Methylglyoxal induces mitochondrial dysfunction and cell death in liver. Toxicol. Res. 2014, 30, 193–198. [Google Scholar] [CrossRef] [PubMed]
  136. Brownlee, M. Biochemistry and molecular cell biology of diabetic complications. Nature 2001, 414, 813–820. [Google Scholar] [CrossRef] [PubMed]
  137. Thornalley, P.J. Glyoxalase I—Structure, function and a critical role in the enzymatic defence against glycation. Biochem. Soc. Trans. 2003, 31, 1343–1348. [Google Scholar] [CrossRef] [PubMed]
  138. Lohmann, K. Beitrag zur enzymatischen Umwandlung von synthetischem Methylglyoxal in Milchsäure. Biochemie 1932, 254, 332–354. [Google Scholar]
  139. Dakin, H.D.; Dudley, H.W. On Glyoxalase. J. Biol. Chem. 1913, 14, 423–431. [Google Scholar]
  140. Neubauer, C. Über die Zerstorung von Milchsäurealdehyd und Methylglyoxal durch tierische Organe. Biochemie 1912, 49, 502–506. [Google Scholar]
  141. Rabbani, N.; Thornalley, P.J. Glyoxalase Centennial conference: Introduction, history of research on the glyoxalase system and future prospects. Biochem. Soc. Trans. 2014, 42, 413–418. [Google Scholar] [CrossRef] [PubMed]
  142. Rabbani, N.; Thornalley, P.J. Dicarbonyls linked to damage in the powerhouse: Glycation of mitochondrial proteins and oxidative stress. Biochem. Soc. Trans. 2008, 36, 1045–1050. [Google Scholar] [CrossRef] [PubMed]
  143. Kanwar, M.; Kowluru, R.A. Role of Glyceraldehyde 3-Phosphate Dehydrogenase in the Development and Progression of Diabetic Retinopathy. Diabetes 2009, 58, 227–234. [Google Scholar] [CrossRef] [PubMed]
  144. Schlotterer, A.; Kukudov, G.; Bozorgmehr, F.; Hutter, H.; Du, X.; Oikonomou, D.; Ibrahim, Y.; Pfisterer, F.; Rabbani, N.; Thornalley, P.; et al. C. elegans as Model for the Study of High Glucose—Mediated Life Span Reduction. Diabetes 2009, 58, 2450–2456. [Google Scholar] [CrossRef] [PubMed]
  145. Bierhaus, A.; Chevion, S.; Chevion, M.; Hofmann, M.; Quehenberger, P.; Illmer, T.; Luther, T.; Berentshtein, E.; Tritschler, H.; Müller, M.; et al. Advanced glycation end product-induced activation of NF-κB is suppressed by α-lipoic acid in cultured endothelial cells. Diabetes 1997, 46, 1481–1490. [Google Scholar] [CrossRef] [PubMed]
  146. Bierhaus, A.; Humpert, P.M.; Morcos, M.; Wendt, T.; Chavakis, T.; Arnold, B.; Stern, D.M.; Nawroth, P.P. Understanding RAGE, the receptor for advanced glycation end products. J. Mol. Med. Berl. Ger. 2005, 83, 876–886. [Google Scholar] [CrossRef] [PubMed]
  147. Lander, H.M.; Tauras, J.M.; Ogiste, J.S.; Hori, O.; Moss, R.A.; Schmidt, A.M. Activation of the receptor for advanced glycation end products triggers a p21(ras)-dependent mitogen-activated protein kinase pathway regulated by oxidant stress. J. Biol. Chem. 1997, 272, 17810–17814. [Google Scholar] [CrossRef] [PubMed]
  148. Thornalley, P.J. Glutathione-dependent detoxification of alpha-oxoaldehydes by the glyoxalase system: Involvement in disease mechanisms and antiproliferative activity of glyoxalase I inhibitors. Chem. Biol. Interact. 1998, 111–112, 137–151. [Google Scholar] [CrossRef]
  149. Bierhaus, A.; Schiekofer, S.; Schwaninger, M.; Andrassy, M.; Humpert, P.M.; Chen, J.; Hong, M.; Luther, T.; Henle, T.; Klöting, I.; et al. Diabetes-associated sustained activation of the transcription factor nuclear factor-kappaB. Diabetes 2001, 50, 2792–2808. [Google Scholar] [CrossRef] [PubMed]
  150. Bierhaus, A.; Stoyanov, S.; Haag, G.; Konrade, I. RAGE-deficiency reduces diabetes-associated impairment of glyoxalase-1 in neuronal cells. Diabetes 2006, 55, A511. [Google Scholar]
  151. Liliensiek, B.; Weigand, M.A.; Bierhaus, A.; Nicklas, W.; Kasper, M.; Hofer, S.; Plachky, J.; Gröne, H.-J.; Kurschus, F.C.; Schmidt, A.M.; et al. Receptor for advanced glycation end products (RAGE) regulates sepsis but not the adaptive immune response. J. Clin. Investig. 2004, 113, 1641–1650. [Google Scholar] [CrossRef] [PubMed]
  152. Hofer, S.; Uhle, F.; Fleming, T.; Hell, C.; Schmoch, T.; Bruckner, T.; Weigand, M.A.; Brenner, T. RAGE-mediated inflammation in patients with septic shock. J. Surg. Res. 2016, 202, 315–327. [Google Scholar] [CrossRef] [PubMed]
  153. Van Herreweghe, F.; Mao, J.; Chaplen, F.W.R.; Grooten, J.; Gevaert, K.; Vandekerckhove, J.; Vancompernolle, K. Tumor necrosis factor-induced modulation of glyoxalase I activities through phosphorylation by PKA results in cell death and is accompanied by the formation of a specific methylglyoxal-derived AGE. Proc. Natl. Acad. Sci. USA 2002, 99, 949–954. [Google Scholar] [CrossRef] [PubMed]
  154. Bourne, H.; Lichtenstein, L.; Melmon, K.; Henney, C.; Weinstein, Y.; Shearer, G. Modulation of inflammation and immunity by cyclic AMP. Science 1974, 184, 19–28. [Google Scholar] [CrossRef] [PubMed]
  155. Janssens, P.A.; Lowrey, P. Hormonal regulation of hepatic glycogenolysis in the carp, Cyprinus carpio. Am. J. Physiol. Regul. Integr. Comp. Physiol. 1987, 252, R653–R660. [Google Scholar]
  156. Beavo, J.A.; Brunton, L.L. Cyclic nucleotide research—Still expanding after half a century. Nat. Rev. Mol. Cell Biol. 2002, 3, 710–718. [Google Scholar] [CrossRef] [PubMed]
  157. Zuo, L.; Shi, L.; Yan, F. The reciprocal interaction of sympathetic nervous system and cAMP-PKA-NF-κB pathway in immune suppression after experimental stroke. Neurosci. Lett. 2016, 627, 205–210. [Google Scholar] [CrossRef] [PubMed]
  158. Park, T.; Chen, H.; Kevala, K.; Lee, J.-W.; Kim, H.-Y. N-Docosahexaenoylethanolamine ameliorates LPS-induced neuroinflammation via cAMP/PKA-dependent signaling. J. Neuroinflamm. 2016, 13, 284. [Google Scholar] [CrossRef] [PubMed]
  159. Rivkin, I.; Neutze, J.A. Influence of cyclic nucleotides and a phosphodiesterase inhibitor on in vitro human blood neutrophil chemotaxis. Arch. Int. Pharmacodyn. Ther. 1977, 228, 196–204. [Google Scholar] [PubMed]
  160. Marone, G.; Columbo, M.; Triggiani, M.; Cirillo, R.; Genovese, A.; Formisano, S. Inhibition of IgE-mediated release of histamine and peptide leukotriene from human basophils and mast cells by forskolin. Biochem. Pharmacol. 1987, 36, 13–20. [Google Scholar] [CrossRef]
  161. Nielson, C.P. Beta-adrenergic modulation of the polymorphonuclear leukocyte respiratory burst is dependent upon the mechanism of cell activation. J. Immunol. 1987, 139, 2392–2397. [Google Scholar] [PubMed]
  162. Dent, G.; Giembycz, M.A.; Evans, P.M.; Rabe, K.F.; Barnes, P.J. Suppression of human eosinophil respiratory burst and cyclic AMP hydrolysis by inhibitors of type IV phosphodiesterase: Interaction with the beta adrenoceptor agonist albuterol. J. Pharmacol. Exp. Ther. 1994, 271, 1167–1174. [Google Scholar] [PubMed]
  163. Lorenowicz, M.J.; Fernandez-Borja, M.; Hordijk, P.L. cAMP signaling in leukocyte transendothelial migration. Arterioscler. Thromb. Vasc. Biol. 2007, 27, 1014–1022. [Google Scholar] [CrossRef] [PubMed]
  164. Skalhegg, B.S.; Tasken, K. Specificity in the cAMP/PKA signaling pathway. Differential expression, regulation, and subcellular localization of subunits of PKA. Front. Biosci. 2000, 5, D678–D693. [Google Scholar] [CrossRef] [PubMed]
  165. Colledge, M.; Scott, J.D. AKAPs: From structure to function. Trends Cell Biol. 1999, 9, 216–221. [Google Scholar] [CrossRef]
  166. Gerlo, S.; Kooijman, R.; Beck, I.M.; Kolmus, K.; Spooren, A.; Haegeman, G. Cyclic AMP: A selective modulator of NF-κB action. Cell. Mol. Life Sci. 2011, 68, 3823–3841. [Google Scholar] [CrossRef] [PubMed]
  167. Morgenstern, J.; Fleming, T.; Schumacher, D.; Eckstein, V.; Freichel, M.; Herzig, S.; Nawroth, P. Loss of glyoxalase 1 induces compensatory mechanism to achieve dicarbonyl detoxification in mammalian Schwann cells. J. Biol. Chem. 2017, 292, 3224–3238. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The “Warburg effect” in activated immune cells. Visualization of the interaction of glycolysis and the tricarboxylic acid cycle (TCA) pathway in naive (left side) and activated (right side) immune cells. While energy production relies on the TCA in native immune cells (solid arrows highlighted in blue) and Glycolysis is of minor importance (dashed arrows left side), the TCA is blocked on two breaking points in activated immune cells, which results in citrate and succinate accumulation. As a result, energy production in activated immune cells relies on Glycolysis (solid arrows highlighted in orange) while the TCA is of minor importance (dashed arrows). Upward arrows symbolize increased levels. Abbreviations: Pyr = Pyruvate, NADH = Nicotinamide adenine dinucleotide (NAD+ in its oxidated form), GDP/GTP = Guanosine diphosphate/Guanosine triphosphate, HIF1α = Hypoxia-inducible factor 1-α, IL-1β = Interleukin 1 β, NO = nitric oxide.
Figure 1. The “Warburg effect” in activated immune cells. Visualization of the interaction of glycolysis and the tricarboxylic acid cycle (TCA) pathway in naive (left side) and activated (right side) immune cells. While energy production relies on the TCA in native immune cells (solid arrows highlighted in blue) and Glycolysis is of minor importance (dashed arrows left side), the TCA is blocked on two breaking points in activated immune cells, which results in citrate and succinate accumulation. As a result, energy production in activated immune cells relies on Glycolysis (solid arrows highlighted in orange) while the TCA is of minor importance (dashed arrows). Upward arrows symbolize increased levels. Abbreviations: Pyr = Pyruvate, NADH = Nicotinamide adenine dinucleotide (NAD+ in its oxidated form), GDP/GTP = Guanosine diphosphate/Guanosine triphosphate, HIF1α = Hypoxia-inducible factor 1-α, IL-1β = Interleukin 1 β, NO = nitric oxide.
Ijms 18 00657 g001
Figure 2. Regulation of MG-derived carbonyl stress in sepsis. Aside from enhanced glycolysis, increased formation of ROS with subsequent inhibition of GAPDH and the glyoxalase system in addition to glutathione consumption contributes to MG-derived carbonyl stress in sepsis. Small upward arrows symbolize increased levels. While solid arrows indicate pathway sequences, dashed arrows indicate reduced steps of a pathway. Abbreviations: GLUT = glucose transporter, MG = Methylglyoxal, MG-AGE = MG derived advanced glycation educts, GA3P = Glyceraldehyde 3-phosphate, DHAP = Dihydroxyacetone phosphate, GAPDH = Glyceraldehyde 3-phosphate dehydrogenase, GLO = Glyoxalase, GSH = Glutathion, TCA = tricarboxylic acid, ROS = reactive oxygen species, NADPH = Nicotinamide adenine dinucleotide phosphate, Nrf2 = Nuclear factor 2, NFκB = Nuclear factor κ B, AP-1 = Activator protein 1, Glo-6-P = Glucose-6-Phosphate, Fru-6-P = Fructose-6-Phosphate, Fru-1,6-bP = Fructose-1,6-Bisphosphate.
Figure 2. Regulation of MG-derived carbonyl stress in sepsis. Aside from enhanced glycolysis, increased formation of ROS with subsequent inhibition of GAPDH and the glyoxalase system in addition to glutathione consumption contributes to MG-derived carbonyl stress in sepsis. Small upward arrows symbolize increased levels. While solid arrows indicate pathway sequences, dashed arrows indicate reduced steps of a pathway. Abbreviations: GLUT = glucose transporter, MG = Methylglyoxal, MG-AGE = MG derived advanced glycation educts, GA3P = Glyceraldehyde 3-phosphate, DHAP = Dihydroxyacetone phosphate, GAPDH = Glyceraldehyde 3-phosphate dehydrogenase, GLO = Glyoxalase, GSH = Glutathion, TCA = tricarboxylic acid, ROS = reactive oxygen species, NADPH = Nicotinamide adenine dinucleotide phosphate, Nrf2 = Nuclear factor 2, NFκB = Nuclear factor κ B, AP-1 = Activator protein 1, Glo-6-P = Glucose-6-Phosphate, Fru-6-P = Fructose-6-Phosphate, Fru-1,6-bP = Fructose-1,6-Bisphosphate.
Ijms 18 00657 g002
Table 1. Overview of the most dominant hormonal changes in systemic inflammation and their contribution to the observed metabolic changes. Upward arrows symbolize “upregualation”, downward arrows symbolize “downregulation”.
Table 1. Overview of the most dominant hormonal changes in systemic inflammation and their contribution to the observed metabolic changes. Upward arrows symbolize “upregualation”, downward arrows symbolize “downregulation”.
HormonePathwayMetabolic Changes
Increased
Cortisol
binding to the glucocorticoid receptor in the nucleus
→ expression of enzymes involved in gluconeogenesis ↑ and β2-adreno receptors ↑ and further anti-inflammatory proteins (e.g., lipocortin, Interleukin-1-Receptor Antagonist (IL-1RA), I κB Kinase)
inhibition of transcription of nuclear factor κ B (NFκB)-dependent genes
gluconeogenesis ↑ in hepatocytes
lipolysis ↑ proteolysis ↑
Nor-/Epinephrine
via β2-receptors (cAMP ↑) in liver and skeletal muscle
via β3-Receptors (cAMP ↑) in adipocytes
lipolysis ↑ and gluconeogenesis ↑ in liver and skeletal muscle.
lipolysis ↑ and ketogenesis ↑ in adipocytes
Vasopressin
via V1-receptors
glycogenolysis ↑
Insulin (although its effects are impaired by peripheral insulin resistance)
binding to the insulin receptor
→ activation and deactivation of enzymes via kinase cascades involving phosphoinositid-3-kinase, the PI3-cascade and the activation of protein kinase B (PKB).
In hepatocytes and skeletal muscle
insertion of glucose transporter type 4 (GLUT4) molecules into the cell membranes of adipocytes and skeletal muscles
glycogen synthesis ↑
gluconeogenesis ↓
triglyceride synthesis ↑
lipolysis ↓
cell growth ↑
cell proliferation ↑
autophagy ↓
Glucagon
binding to G protein-coupled receptors−
→ cAMP ↑ → protein kinase A activity ↑
glycogenolysis ↑
Reduced
Thyroid-stimulating hormone (TSH) ↓, triiodothyronine (T3) ↓, thyroxine (T4) ↓
changes of gene expression
insertion of β-adreno receptors in cell membranes ↓
insertion of α-adreno receptors in cell membranes ↑
insulin secretion ↓
corticosteroid and catecholamine production and secretion ↓

Share and Cite

MDPI and ACS Style

Schmoch, T.; Uhle, F.; Siegler, B.H.; Fleming, T.; Morgenstern, J.; Nawroth, P.P.; Weigand, M.A.; Brenner, T. The Glyoxalase System and Methylglyoxal-Derived Carbonyl Stress in Sepsis: Glycotoxic Aspects of Sepsis Pathophysiology. Int. J. Mol. Sci. 2017, 18, 657. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms18030657

AMA Style

Schmoch T, Uhle F, Siegler BH, Fleming T, Morgenstern J, Nawroth PP, Weigand MA, Brenner T. The Glyoxalase System and Methylglyoxal-Derived Carbonyl Stress in Sepsis: Glycotoxic Aspects of Sepsis Pathophysiology. International Journal of Molecular Sciences. 2017; 18(3):657. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms18030657

Chicago/Turabian Style

Schmoch, Thomas, Florian Uhle, Benedikt H. Siegler, Thomas Fleming, Jakob Morgenstern, Peter P. Nawroth, Markus A. Weigand, and Thorsten Brenner. 2017. "The Glyoxalase System and Methylglyoxal-Derived Carbonyl Stress in Sepsis: Glycotoxic Aspects of Sepsis Pathophysiology" International Journal of Molecular Sciences 18, no. 3: 657. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms18030657

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop