Abstract
Cellular proteins are decorated with a wide range of acetyl and other acyl modifications. Many studies have demonstrated regulation of site-specific acetylation by acetyltransferases and deacetylases. Acylation is emerging as a new type of lysine modification, but less is known about its overall regulatory role. Furthermore, the mechanisms of lysine acylation, its overlap with protein acetylation, and how it influences cellular function are major unanswered questions in the field. In this review, we discuss the known roles of acetyltransferases and deacetylases, and the sirtuins as a conserved family of NAD+-dependent protein deacylases that are important for response to cellular stress and homeostasis. We also consider the evidence for an emerging idea of non-enzymatic protein acylation. Finally, we put forward the hypothesis that protein acylation is a form of protein “carbon stress”, that the deacylases evolved to remove as a part of a global protein quality control network.
Introduction
Acetylation is an evolutionarily conserved post-translational modification occurring on lysine residues. Using acetyl-CoA as a donor, acetyltransferase enzymes catalyze site-specific acetylation of several proteins; a large family of deacetylase enzymes opposes this reaction by removing acetylation. This balanced regulation was first discovered on histones (Allfrey et al., 1964; Isenberg, 1979; Strahl and Allis, 2000), but has since been shown to regulate several biological processes, including gene and protein expression, cell survival, and cell death (Choudhary et al., 2009; Verma et al., 2012; Yi et al., 2011).
Sirtuins are one family of deacetylases that plays a role in a variety of important homeostatic and stress-related cellular processes. They have a unique ability to remove acetyl and other acyl groups from protein lysine side chains in a nicotinamide adenine dinucleotide (NAD+)-dependent manner (Imai et al., 2000). The founding member of the Sirtuin family, yeast Sir2p, was originally identified in a genetic screen for gene silencing factors (Rine et al., 1979). Interest in the sirtuin family was stimulated by the findings that yeast lifespan is extended by increasing Sir2 gene dosage and that Sir2 utilizes the metabolic coenzyme NAD+ to deacetylate its targets, thereby providing a link between metabolism, gene expression, and lifespan (Imai et al., 2000; Kaeberlein et al., 1999).
Mammalian sirtuins are grouped into four phylogenetic classes: Class I contains SIRT1-3, class II contains SIRT4, class III contains SIRT5, and class IV contains SIRT6 and SIRT7 (Frye, 2000). All mammalian sirtuins were originally thought to have deacetylase activity owing to their homology to yeast Sir2. Indeed, class I sirtuins (SIRT1-3) have robust deacetylase activity on a wide variety of substrates (Michishita et al., 2005). Because the remaining classes of sirtuins (SIRT4-7) have low deacetylase activity on tested substrates, they were thought to be either highly specific deacetylases or ADP-ribosyltransferases. Recently, SIRT5 was shown to have robust lysine desuccinylase and demalonylase activities, expanding the known enzymatic activities of the sirtuin family (Du et al., 2011; Peng et al., 2011). These findings suggested that the different sirtuins in different phylogenetic classes could harbor not-yet-discovered enzymatic activities (Hirschey, 2011). This idea is supported by recent studies demonstrating that SIRT6 has efficient lysine depalmitoylase and demyristoylase activity (Jiang et al., 2013). Another recent study describes deglutarylase activity for SIRT5 (Tan, 2014). Remarkably, another study showed in vitro that many of the sirtuins are capable of hydrolyzing long-chain acyl, including octanoyl-, decanoyl-, and dodecanoyl-lysines (Feldman et al., 2013). However, the extent to which sirtuins are removing these longer-chain acyl modifications in vivo has not yet been explored. Together, these studies show that mammalian sirtuins are not merely deacetylases, but are NAD+-dependent deacylases capable of acting on a several acyl substrates.
Sirtuins have important roles in a wide range of mammalian biological processes, including metabolic homeostasis (e.g. fasting, CR, high-fat diet feeding), and resistance to stresses (e.g. DNA damage, ROS, protein misfolding) (Cheng et al., 2003; Finley et al., 2011; Hirschey et al., 2011; Toiber et al., 2013; Westerheide et al., 2009). Oftentimes sirtuin-mediated deacylation is associated favorable cellular processes (Hall et al., 2013), and models of sirtuin ablation or knock-down are associated with increased incidence or acceleration of disease (Finkel et al., 2009). However, sirtuin ablation is not always pathogenic. For example, mice lacking SIRT2 are markedly protected from ischemic injury (Narayan et al., 2012). Additionally, SIRT1 has been described to both promote and suppress tumorigenesis, depending on the molecular and cellular context (Yuan et al., 2013). However, the mammalian sirtuins are collectively viewed to increase stress resistance by remodeling metabolism, altering inflammatory processes, and enhancing oxidative stress responses (Hall et al., 2013). Yet, how and why removing an acyl-group from a protein is generally associated these processes is unclear.
Despite the number of studies on the sirtuins and their function, less is known about the family of acetyltransferase or acyltransferase enzymes opposing sirtuin enzymatic activity. Lysine acetyltransferases such as those in the GNAT, MYST, and P300/CBP superfamilies have been well-studied in the context of chromatin regulation, but also regulate the stability, localization, and interactions of several extra-nuclear protein targets (Jiang et al., 2011; Lin et al., 2013; Zhao et al., 2013); for a recent review, see (Friedmann and Marmorstein, 2013). Remarkably, none of these acetyltransferases have known roles in the mitochondria where protein acylation is highly abundant. In this compartment, little is known about the mechanism of acylation. One report described a component of an acetyltransferase, or the acetyltransferase machinery, in the mitochondria called GCN5L1 (Scott et al., 2012). Another recent report described an acetyltransferase activity for the mitochondrial enzyme ACAT1 (Fan et al., 2014). However, more studies on the enzymology of these reactions are required to determine how these proteins catalyze acetyl transfer to proteins.
In contrast, recent and historical evidence supports the idea that non-enzymatic acetylation and acylation can lead to protein modification (Wagner and Payne, 2013; Weinert et al., 2013b). Indeed, the idea of non-enzymatic mechanisms is emerging as a possible contributor to the protein acetylation and acylation landscape (Guan and Xiong, 2011; Newman et al., 2012). In this review, we analyze the evidence for non-enzymatic protein acetylation and acylation, and discuss the ability of the sirtuins to regulate these nonenzymatic processes. Finally, we propose a model where the sirtuins are regulators of nonenzymatic acylation, and are therefore important mediators of overall protein quality control (Figure 1).
Figure 1. Model of sirtuin-mediated protein quality control.
(A) Intrinsically reactive metabolites can non-enzymatically react with nucleophilic protein residues to form post-translational modifications which can compromise protein function and could be considered “carbon stress”. (B) Non-enzymatic acyl-modifications can be targeted for removal by the sirtuin family of NAD+-dependent lysine deacylases to restore protein function and cellular health, which is part of a global protein quality control system.
Non-enzymatic Protein Acetylation
The use and interconversion of carbon substrates is a near universal property of what defines life. Within this simplistic concept of cellular metabolism, chemical reactivity is a fundamental and often overlooked requirement for metabolism to operate. Whether metabolism is pro-growth (anabolic), pro-breakdown (catabolic), or both (amphibolic), metabolites and metabolic intermediates must be chemically reactive for physiological use.
The reactivity of a cellular metabolite is often dependent on the presence or absence of nucleophilic or electrophilic moieties and/or leaving groups. For example, the carbonyl group is electrophilic and reactive, and certain metabolic enzymes can further “activate” carbonyl-containing compounds by covalently adding additional electronegative functional groups, such as thiols or phosphates. These additional groups are also known as leaving groups for their tendency to “leave” the compound after a nucleophilic attack, and the inherent reactivity of a given chemical or metabolite can often be determined by whether or not it contains a “good” leaving group (Figure 2A).
Figure 2. The reactivity of functional groups in biology.
(A) (Left) Carbonyl-containing functional groups commonly used in biological systems are ranked according to their inherent reactivity. (Right) Specific examples of cellular metabolites containing the select carbonyl compounds are listed along with the biological processes in which they are generated or consumed. Carbonyl groups are highlighted in blue and leaving groups are highlighted in red. (B) Reactive acyl-containing metabolites and the mechanisms leading to non-enzymatic protein modification. AGE, advanced glycation end products.
Compounds containing a reactive thioester group in the form of coenzyme A (CoA) are particularly prevalent throughout metabolism in diverse organisms. The activation of metabolites with a labile thioester bond facilitates many biochemical processes including de novo lipogenesis, the TCA cycle, ketogenesis, amino acid metabolism, as well as protein lysine acetylation. The chemical versatility of the thioester bond inspired the “Thioester World” hypothesis, which suggests a central role of this chemical group in the process of abiogenesis and early chemical evolution (De Duve, 1995).
While organisms have evolved to rely on intrinsically reactive metabolites to carry out their vital cellular functions, the utility of reactive metabolites in metabolism also poses a challenge: reactive metabolic intermediates could spontaneously react with nearby proteins. For example, acetylation in bacteria correlates strongly with fluctuating levels of the highly reactive intermediate acetyl-phosphate (Figure 2A) (Weinert et al., 2013b). In this study, the majority of protein lysine acetylation in E. coli occurs independently of its only known acetyltransferase, YfiQ. Furthermore, incubating bacterial proteins with acetyl-phosphate in vitro causes acetylation at nearly all the sites identified in vivo, and approximately 10% of the putatively non-enzymatic acetylation events are targeted by the sirtuin deacetylase CobB (Weinert et al., 2013b). This study provides strong evidence that acetyl-phosphate mediates global shifts in protein acetylation via non-enzymatic mechanisms, and that the sirtuin CobB regulates only a subset of these chemical acetylation events.
Non-enzymatic protein lysine acetylation mediated by reactive acetyl-thiol compounds (i.e. acetyl-CoA) has been reported on human serum albumin, purified histones, lysine-containing polypeptides, and recombinant bacterial proteins (Baddiley et al., 1952; d’Alayer et al., 2007; Paik et al., 1970). In fact, acetyl-CoA has historically been described as a suitable “acetylating agent” (Delpech et al., 1983; Paik et al., 1970). More recently, nonenzymatic reactions between acetyl-CoA and protein lysines have been shown and discussed in context of their ability to contribute to the acetylation landscape in biological systems.
For example, one study showed that while the acetyltransferase GCN5 can catalyze acetylation of proteins, neither the enzyme nor its activity is required for acetylation in some conditions (Tanner et al., 2000; Tanner et al., 1999). Specifically, the catalytic mechanism of the GCN5 acetyltransferase involves deprotonating the epsilon amino group of a lysine residue, which enables it to act as a nucleophile towards the electrophilic carbonyl carbon in acetyl-CoA, resulting in an acetylated lysine. Because deprotonation of lysine residues is pH-dependent, acetyl-CoA incubation with catalytically-dead GCN5 or no-enzyme control reactions performed at alkaline pH still leads to lysine acetylation, primarily by increasing the number of amino groups acting as nucleophiles towards the inherently reactive acetyl-CoA. These data suggest that some cytoplasmic and/or nuclear protein acetylation could occur without its acetyltransferase.
As another example, the environment of the mitochondrial matrix, which has an alkaline pH and abundant acetyl-CoA, could make non-enzymatic acetylation favorable (Casey et al., 2010; Garland et al., 1965). One recent study demonstrated that reproducing these conditions in vitro is sufficient to cause non-enzymatic acetylation of mitochondrial and non-mitochondrial proteins using acetyl-CoA as a substrate (Wagner and Payne, 2013). An additional study showed that acetylation accumulates in growth-arrested Saccharomyces cerevisiae that depended on acetyl-CoA generation in distinct subcellular compartments (Weinert et al., 2013a), suggesting that non-enzymatic protein acetylation contributes to the overall acetylation landscape.
Further supporting this idea, recent findings from eukaryotic systems indicate that the stoichiometry of acetyl-lysine modifications is low. For example, 86% of acetylation sites identified in yeast are estimated to have less than 1% stoichiometry (Weinert et al., 2013a). These findings are in contrast to the best measurements of stoichiometry of phosphorylation in yeast, where only 11% of phosphorylation sites show less than 1% stoichiometry (Wu et al., 2011). Interestingly, the most abundant (high stoichiometric) acetylation sites in yeast occurred on histones, proteins present in histone deacetylase (HDAC) or histone acetyltransferase (HAT) complexes, and on transcription factors, supporting the idea that acetylation can be regulated in these complexes (Weinert et al., 2013a). In contrast, baseline mitochondrial protein acetylation levels are low in yeast, and increase markedly upon increasing acetyl-CoA concentrations in vivo by manipulating acetyl-CoA consuming enzymes (citrate synthase), or were reduced when an acetyl-CoA generating enzyme was ablated (pyruvate dehydrogenase). These studies indicate that a large portion of protein lysine acylation lacks the high levels of stoichiometry typically associated with regulatory signaling modifications such as phosphorylation, supporting the idea that some protein acetylation could occur in a non-targeted, non-signaling manner (i.e. non-enzymatically).
An important caveat is that these studies were performed in yeast, and the stoichiometry of acetylation was only estimated indirectly. Therefore, absolute levels of mammalian protein acetylation are not known. The major reason for this lack of understanding is because measuring the stoichiometry of global acetylation sites accurately would require thousands of isotopically labeled, custom-synthesized, internal standards and is therefore impractical with current technology. Thus, studies will generally determine “maximal median stoichiometry” by normalizing acetyl-lysine ratios to protein ratios and reporting the findings as relative values (Hebert et al., 2013; Rardin et al., 2013b). Using these techniques, low overall stoichiometry of protein acetylation is still observed.
Together, these studies support the notion that non-enzymatic protein acetylation by acetyl-CoA could occur, and would be especially favorable in the mitochondria. Importantly, low-level acetylation mediated by a promiscuous acetyltransferase that is sensitive to changes in acetyl-CoA levels cannot be formally excluded. However, the historical ability of acetyl-CoA to acetylate a wide variety of substrates, coupled with recent evidence of the dependence on the generation of distinct acetyl-CoA pools for acetylation, are highly consistent with a non-enzymatic mechanism contributing to the global protein acetylation landscape.
Non-enzymatic Protein Acylation
This model further suggests that non-enzymatic acetylation might not be limited to acetyl-CoA alone. Like acetyl-CoA, succinyl-CoA and malonyl-CoA contain a thioester bond which is intrinsically reactive toward nucleophiles (Figure 2A). Indeed, protein lysine succinylation and malonylation was found on a wide-range of mitochondrial, cytosolic, and nuclear proteins (Du et al., 2011; Peng et al., 2011; Zhang et al., 2010).
The reactive nature of succinyl-CoA suggests that manipulating concentrations of this intermediate in vivo might alter protein succinylation. Indeed, historical studies demonstrated non-enzymatic protein succinylation could be regulated by succinyl-CoA concentrations and was a potential control mechanism for ketogenesis (Lowe and Tubbs, 1985; Quant et al., 1990). More recently, succinyl-CoA has been shown to non-enzymatically succinylate protein lysine residues in a pH-dependent manner at micromolar concentrations (Wagner and Payne, 2013; Weinert et al., 2013c). Furthermore, deleting succinyl-CoA generating and consuming enzymes in the TCA cycle, in yeast, caused significant global decreases and increases in succinylation, respectively (Weinert et al., 2013c).
Similarly, a new protein modification called lysine glutarylation is enriched on mitochondrial proteins (Tan, 2014). Like acetyl-CoA and succinyl-CoA, glutaryl-CoA is primarily generated in mitochondria and protein glutarylation is found on a wide-range of metabolic proteins. In vitro, incubation of heat-inactivated mitochondrial proteins with micromolar concentrations of glutaryl-CoA, but not glutarate, increases protein glutarylation. In vivo, glutarylation increases when flies are fed glutaryl-CoA precursors or when protein catabolism is increased.
The prevalence of both lysine acetylation and malonylation/succinylation/glutarylation on proteins involved in mitochondrial metabolism suggests something unique about this cellular compartment. Furthermore, similar to lysine acetylation, a significant proportion of succinylated proteins are localized to mitochondria, with less succinylation among cytosolic and nuclear proteins (Park et al., 2013). Indeed, the mitochondrial matrix is the primary site of acyl-CoA metabolism, where acetyl-CoA and succinyl-CoA intermediates reach estimated steady-state concentrations in the millimolar range (Garland et al., 1965; Hansford and Johnson, 1975).
Interestingly, early estimates on the levels and stoichiometry of succinylation have shown 68% of succinylation sites quantified from a subset of proteins in mouse embryonic fibroblasts display less than 10% stoichiometry (Park et al., 2013). These results are in agreement with the low stoichiometry observations of acetylation, described above. Additionally, despite the different chemical properties of acetyl-lysine and succinyl-lysine modifications, significant overlap is found between the protein lysine residues that are acetylated and succinylated (Rardin et al., 2013a; Weinert et al., 2013c), supporting the idea that some lysines are susceptible to non-enzymatic acylation by reactive acyl-CoAs.
Lysine malonyltransferase, succinyltransferase, or glutaryltransferase enzymes have not been identified in any subcellular compartment (Park et al., 2013; Weinert et al., 2013c). In the mitochondria, the biochemical properties of the protein microenvironment, such as metabolite concentration and pH, may be facilitating non-enzymatic acylation reactions. While enzymatic acyltransfer cannot be ruled out, acyl-CoA concentrations in the mitochondria are several hundred-fold higher than the kinetic binding constants (Km) of characterized lysine acetyltransferases (Albaugh et al., 2011). Thus, if an acyltransferase resides in the mitochondria that is kinetically comparable to known acetyltransferases, its catalytic activity would be fully saturated at all times. This would either lead to high stoichiometric levels of acylation (which has not yet been observed), or would leave a substantial excess of reactive acetyl-CoA in a steady state to potentially react with protein amine groups.
Together, these studies suggest that non-enzymatic protein lysine acylation may occur in diverse organisms and is closely correlated with perturbations of acyl-CoA metabolism. Furthermore, protein acylation would be particularly favorable in the mitochondrial compartment. Studies in this area so far suggest that non-enzymatic reactions between metabolites and proteins might be a common mechanism of protein acylation.
Protein Acylation as Carbon Stress
One of the major challenges of studying protein acetylation and acylation is understanding the overall function of these post-translational modifications. Indeed, the seemingly infinite number of possible combinations of modifications, on a given protein, amplified across the proteome is daunting. So far, investigations on acylation have either quantified the acyl proteome using mass spectrometry-based proteomics, taken a one-target-at-a-time approach, or both. Large-scale proteomic surveys have identified metabolic proteins are over-represented in acetylation surveys, suggesting that acetylation could regulate cellular metabolism. Acetylation of metabolic proteins outside the mitochondria can be activating, inhibitory, or have no measureable consequence. However, one theme that is emerging across mitochondrial metabolic studies is that the effect of protein acetylation is consistently inhibitory, especially to mitochondrial energy-producing processes (Ghanta et al., 2013). Interestingly, protein acetylation and acylation have significant overlap in the mitochondria (Rardin et al., 2013a). Together, these observations support the idea that protein acylation in the mitochondria is largely inhibitory.
Based on a non-enzymatic mechanism of action, we propose that protein acylation is one form of “carbon stress,” in which the accumulation of intrinsically reactive carbon metabolites can directly and negatively alter protein function and disrupt cellular homeostasis. Using this definition, other endogenous examples of carbon stress include protein glycation, carbamylation (Figure 2B), and the recently identified 3-phosphoglycerylation (Moellering and Cravatt, 2013). All of these modifications can be driven by non-enzymatic reactions of metabolites with proteins. Under basal conditions, these modifications might occur at a low level. During physiological or pathological states that increase the amounts of reactive metabolite, non-enzymatic reactions could also increase and have a detrimental effect on protein function; in this setting, these reactions would be considered a form of “carbon stress.”
Acylation could alter protein function by a variety of mechanisms. For example, lysine acetylation neutralizes a positive charge and alters hydrogen bonding potential, which could disrupt protein-protein or protein-subunit interactions, enzyme-substrate and cofactor binding, and allosteric regulation. Some of these mechanisms have been demonstrated experimentally (Bharathi et al., 2013), while others are predicted through protein structural analyses (Hebert et al., 2013; Shimazu et al., 2010; Tao et al., 2010). Together, these studies support the idea that unregulated protein acylation could broadly interfere with protein function and compromise cellular health, potentially leading to disease.
Carbon Stress and Sirtuin-Mediated Protein Quality Control
Sirtuins are a conserved family of NAD+-dependent protein deacylases that are important in response to cellular stress and homeostasis. Mounting evidence shows that the sirtuins act on a wide variety of substrates. While the role of sirtuins is complex, and can be unique in certain tissues or organ systems, sirtuin-mediated deacylation is often associated with increased protein function and metabolic health; for a review, see (Finkel et al., 2009).
Indeed, increases in mitochondrial protein acetylation (by loss or ablation of SIRT3) leads to acceleration and/or development of several disease states, including cardiac hypertrophy, hearing loss, the metabolic syndrome, and reduced regenerative capacity of hematopoietic stem cells (Brown et al., 2013; Hafner et al., 2010; Hirschey et al., 2011; Someya et al., 2010). Furthermore, hypersuccinylation (by ablation of SIRT5) of numerous mitochondrial and non-mitochondrial proteins (Park et al., 2013; Yu et al., 2013) leads to hyperammonemia (Nakagawa et al., 2009) and reduced fatty acid oxidation and less ketogenesis (Rardin et al., 2013a). Finally, protein hyperglutarylation (by ablation of SIRT5) is also associated with reduced carbamoyl phosphate synthase 1 activity (Tan, 2014). Thus, protein acylation may be a form of non-enzymatic chemical damage which alters protein function and compromises cellular health.
Interestingly, studies investigating the effects of acetylation on specific SIRT3 target enzymes indicate that many known acetylation events do not affect enzymatic activity (Bharathi et al., 2013; Hallows et al., 2006; Hirschey et al., 2010; Shimazu et al., 2010; Tao et al., 2010; Yu et al., 2012). Large-scale proteomic surveys support this idea, and have shown that the overall amount of acetylation predicted to be regulated by SIRT3 in the mitochondria is low (~20%) compared to the majority of mitochondrial protein acetylation, which remains relatively static (Hebert et al., 2013). These findings support a model in which a subset of protein acetylation has consequences for protein function, and are regulated by SIRT3, whereas others sites of protein acetylation have no relevance on protein function. Interestingly, this model is analogous to how the bacterial sirtuin CobB regulates protein acetylation during stationary phase growth: when concentrations of reactive acetyl-phosphate and global protein acetylation concurrently increase, CobB removes a subset of protein acetylation (Figures 3A and 3B) (Weinert et al., 2013b).
Figure 3. Sirtuin-mediated deacetylation is analogous to other protein quality control mechanisms.
(A) Reactive acetyl-CoA is predicted to increase in mice undergoing calorie restriction, which is correlated with hyperacetylation of mitochondrial proteins. In this setting, the mitochondrial deacetylase SIRT3 maintains the acetylation status of select protein lysine residues at low levels, thereby maintaining protein function and metabolic health (Hebert et al., 2013). (B) During the stationary phase in bacteria, the concentration of reactive acetyl-phosphate increases which is correlated with hyperacetylation of proteins. The bacterial sirtuin deacetylase CobB also regulates the acetylation status of select lysines residues at low levels, which likely serves to maintain protein function and metabolic homeostasis (Weinert et al., 2013b). In an analogous manner, oxidative stress and nitrosative stress cause the oxidation (C) and nitrosylation (D) of protein residues, respectively. These damaging non-enzymatic protein modifications are removed by methionine sulfoxide reductases (MsrA,B) and thioredoxins (Trx) or the S-nitrosoglutathione reductase (GSNOR) system, which preserve protein function and support cellular health.
Importantly, this does not imply that only acylation events targeted by sirtuins have functional consequences. Indeed, some acylation events could occur on catalytic lysines, which are not accessible to or targeted by sirtuin deacylation, but would have an irreversible negative effect on enzymatic activity. However, examples of these types of acylation (i.e. non-sirtuin targeted but functionally relevant) have not yet been described in the literature.
Together, these studies suggest that if non-enzymatic acylation is a contributor to mitochondrial protein acylation and a subset of these modifications are detrimental to protein function, then this form of stress would require a protein quality control response to remove spurious chemical modifications on proteins; indeed, the sirtuins could serve this exact role.
Supporting this idea, some sirtuins might respond to or be upregulated during carbon stress when protein acylation increases. During calorie restriction in mice, global hepatic mitochondrial protein acetylation increases while whole tissue acetyl-CoA levels fall, suggesting increased utilization of acetyl-CoA in mitochondria for energy requiring processes (Hebert et al., 2013). Interestingly, SIRT3 expression increases during calorie restriction, suggesting that SIRT3 may be responding to a condition when increased non-enzymatic protein acetylation would be expected due to an increased mitochondrial acetyl-CoA burden. Mice fed a high-fat diet for one week also show increased SIRT3 expression, further suggesting that SIRT3 is responding to conditions predicted to generate more mitochondrial acetyl-CoA (Hirschey et al., 2011).
While mitochondrial SIRT3 and SIRT5 are well-positioned to play a role in managing non-enzymatic lysine acylation, one of the major questions remaining is whether these proposed roles are broadly applicable to the rest of the sirtuin family. Acetyl-CoA and succinyl-CoA are particularly abundant metabolic intermediates; however there are many other acyl-CoA intermediates utilized throughout cellular metabolism, all of which contain the inherently reactive thioester bond and many of which function outside the mitochondrial compartment (Figure 2). For example, many other endogenous acyl-lysine modifications have been validated, including propionylation, butyrylation, crotonylation, and glutarylation. Importantly, these modifications can be removed by HDACs and sirtuins in vitro (Feldman et al., 2013; Madsen and Olsen, 2012). Additionally, longer chain acyl-lysine modifications occur in cells and are targeted for removal by SIRT6 in vivo (Jiang et al., 2013).
While no dedicated lysine propionyltransferase or butyryltransferase has been identified, some acetyltransferases can catalyze propionylation and butyrylation of histones at substantially lower efficiency than acetylation (Albaugh et al., 2011). Importantly, the influence of potential unidentified lysine acyltransferase enzymes cannot be formally excluded. However, the common reactive nature of acyl-CoA intermediates, coupled with recent findings indicating that sirtuins are capable of removing a broad range of acyl-lysine modifications, provides the rationale for the sirtuins to serve a role to remove protein acylation, reduce carbon stress, and maintain overall protein quality control.
Analogous Stress and Protein Quality Control Systems
Several analogous stressors and protein quality control systems are found throughout biology. Organisms experience several types of physiological stress and have evolved systems to maintain protein quality control and overall protein function. Oxidative stress is a quintessential example, which is generated from reactive oxygen species (ROS). For aerobic organisms, ROS are an unavoidable consequence of using molecular oxygen as the terminal election acceptor during cellular respiration (Davies, 2000). Cellular ROS, such as superoxide and hydroxyradicals, can spontaneously react with and damage cellular macromolecules including DNA, lipids, and proteins (Figure 3C). Biological systems have evolved a suite of quality control strategies to mitigate the harmful effects of ROS. For example, catalase, superoxide dismutases, peroxiredoxins, thioredoxins, glutaredoxins, glutathione, glutathione peroxidases, and methionine sulfoxide reductases all represent families of enzymes which have evolved to directly or indirectly cope with harmful ROS and thereby mitigate potentially damaging non-enzymatic modification to macromolecules (Meyer et al., 2009; Muller et al., 2007).
The importance of these endogenous antioxidant systems in mammalian organisms is demonstrated by the pathological phenotypes elicited by genetically ablating their individual components. For instance, mice lacking the ability to detoxify mitochondrial superoxide due to homozygous deletion of manganese superoxide dismutase (MnSOD or SOD2) develop cardiomyopathy, acidosis, and are neonatal lethal (Li et al., 1995). An inability to detoxify lipid peroxides via germline ablation of glutathione peroxidase 4 (GPX4) results in embryonic lethality, while inducible deletion of GPX4 in adult mice causes rapid degeneration and death (Seiler et al., 2008). Furthermore, genetic ablation of mitochondrial methionine sulfoxide reductase (MsrA) in mice causes an increased sensitivity to oxidative stress, increased protein carbonyl adducts, and a 40% decrease in lifespan (Moskovitz et al., 2001). Moreover, impairments or alterations in many of these antioxidant systems are associated with tumorigenesis in mice and humans (De Luca et al., 2010; Muller et al., 2007). Indeed, oxidative stress and reduced protein quality control have been implicated in many diseases.
Remarkably, its counterpart called “reductive stress” also contributes to reduced protein quality control. Reductive stress is an abnormal increase of reducing equivalents (e.g. reduced glutathione [GSH] or NADPH) that contributes to the molecular pathogenesis of disease (Rajasekaran et al., 2007). In one example, a transgenic mouse model that mimics a human cardiomyopathy caused by a mutation in the gene encoding αB-crystallin (R120GCryAB) had enhanced generation of the reducing equivalents NADPH and GSH, increased protein aggregation, and developed cardiomyopathy.
Additionally, nitrosative stress is a unique type of redox stress induced by nitric oxide or NO-related species (e.g. including S-nitrosothiols) which lowers intracellular thiol levels (Figure 3D) (Hausladen et al., 1996). One cellular and genetic response to this stress is the direct activation of transcription, which constitutes an adaptation to nitrosative stress.
Thus, organisms that experience oxidative, reductive, and nitrosative stresses, and perhaps carbon stress, all require mechanisms to counteract these ongoing stressors. Several systems have evolved to do so, and genetic ablation of these quality control pathways reduces the ability to respond to stress and overall organismal fitness; remarkably, ablation of the sirtuin genes is associated with reduced ability to respond to stress and reduced overall organismal fitness.
Role of Carbon Stress in Disease
Increased carbon stress and/or dysregulation of a carbon stress response system could be a general feature of disease, with reduced protein quality control and therefore reduced protein function as a common underlying mechanism. Indeed, several examples of endogenous carbon metabolites undergoing non-enzymatic reactions with protein amine groups contribute to disease. In the setting of diabetes mellitus, prolonged elevation of aldehyde-containing sugars in the blood and intracellular dicarbonyl compounds cause excessive non-enzymatic glycation of protein lysine and arginine residues which can further react to form advanced glycation end products (AGEs) (Figure 2B) (Thornalley, 2005). Non-enzymatic protein glycation and AGE formation contribute to the pathophysiology of diabetes and are associated with cardiovascular and retinal complications by interfering with protein function and cellular signaling (Brownlee, 2001; Fukami et al., 2013).
Remarkably, Louis-Camille Maillard, who first reported the protein glycation process over 100 years ago using an in vitro system, predicted that this reaction may be relevant in diabetes (Maillard, 1912). However, it was not until the 1980s that the role of endogenous non-enzymatic protein glycation and AGEs in mediating diabetic complications was recognized (Garlick et al., 1984; Ulrich and Cerami, 2001). Interestingly, metabolism of glycation-inducing sugars is predicted to be a key evolutionary driver for the Glyoxylase I and II enzyme system, which serves to detoxify intracellular glycation-inducing dicarbonyl compounds such as methylglyoxal (Rabbani and Thornalley, 2012). Thus, aldehyde-containing sugars are carbon sources essential for viability of many cell-types, but they and their byproducts can react with and modify cellular proteins and therefore require coordinate detoxification mechanisms.
In addition to sugar metabolism, lipid metabolism can lead to protein modification. For example, oxidation of polyunsaturated fatty acids generates lipid aldehydes such as malondialdehyde and 4-hydroxy-2-nonenal which are capable of forming harmful adducts with protein lysine, cysteine, and histidine residues (Dalle-Donne et al., 2003). Furthermore, the metabolism of ethanol produces the toxic reactive intermediate acetaldehyde which can similarly form harmful adducts with protein or DNA (Langevin et al., 2011). The ongoing formation of toxic aldehyde species during metabolism could have provided the evolutionary impetus for the aldehyde dehydrogenase superfamily of enzymes, which together can detoxify both lipid and ethanol-derived aldehyde-containing metabolites (Koppaka et al., 2012). The importance of aldehyde detoxification in mammals is evinced by studies of the aldehyde dehydrogenase 2 (ALDH2) knockout mouse or humans carrying the near inactive variant of ALDH2 (ALDH2*2), which demonstrate severe ethanol sensitivity and an increased risk of several cancers (Yokoyama et al., 1998). Furthermore, excessive protein aldehyde adducts are associated with many human diseases including Alzheimer’s disease, Parkinson’s disease, and diabetes (Dalle-Donne et al., 2003).
Other non-enzymatic reactions involving metabolites are implicated in disease but lack known quality control or detoxifying mechanisms. Similar to the mechanism of protein glycation in diabetes, prolonged elevation of cyanate in the blood of smokers or during renal failure can cause harmful non-enzymatic protein lysine carbamylation (Figure 2B) (Fluckiger et al., 1981; Stark, 1965). Excessive protein carbamylation may be a common mechanism underlying the pathophysiology of inflammation and coronary artery disease (Wang et al., 2007).
Separately, recent work has demonstrated that the glycolytic metabolite and acyl-phosphate 1,3-bisphosphoglycerate (Figure 2) can non-enzymatically modify protein lysine residues by virtue of its highly reactive electrophilic carbonyl carbon (Moellering and Cravatt, 2013). This novel modification, 3-phosphoglyceryl-lysine, was further shown to impair the activity of select glycolytic enzymes, but is not yet associated with any disease states.
More commonly, drug toxicity is often caused by electrophilic drug metabolites nonenzymatically reacting with cellular proteins, and minimizing this problem is one of the foremost challenges during the development and optimization of small-molecule pharmaceuticals (Evans et al., 2004; Horng and Benet, 2013). Interestingly, the reactivity and toxicity of many drug metabolites is dependent on their bioactivation to acyl-CoA thioesters (Sallustio et al., 2000). Collectively, these studies demonstrate that non-enzymatic reactions of endogenously-generated metabolites with cellular proteins contributes to the pathophysiology of diabetes, neurodegeneration, kidney failure, inflammation, coronary artery disease, and drug toxicity.
Carbon Stress, Protein Quality Control, and Aging
Many of the diseases associated with carbon stress and reduced protein quality control (diabetes, cardiovascular disease, neurodegeneration, cancer; described above) are also associated with aging. Sirtuins are also implicated in regulating diseases of aging, as well as in mediating the salutary effects of caloric restriction (Finkel et al., 2009). When sirtuins are examined from the perspective of protein quality control, perhaps it is not surprising that sirtuin-mediated deacylation of cellular proteins has generally protective benefits for cellular and organismal health. Just as genetic loss of conserved protein quality control mechanisms accelerates the development of age-related diseases and/or shortens lifespan in models systems, so does the loss of many sirtuin family members. Conversely, pharmacologic or genetic stimulation of protein quality control machinery can extend lifespan in diverse organisms (Powers et al., 2009).
Recent evidence indicates that a stoichiometric imbalance between mitochondrial genome-encoded and nuclear genome-encoded subunits of the respiratory chain induces mitochondrial protein quality control machinery, which extends lifespan in C. elegans through a hormetic effect and is correlated with lifespan in mice (Houtkooper et al., 2013). Interestingly, further evidence indicates that the longevity-promoting effects of mitochondrial protein quality control are stimulated by SIRT1 and by raising cellular NAD+ levels, suggesting that in addition to the proposed direct effects of sirtuin-mediated deacylation on protein quality control, sirtuins also indirectly regulate the protein quality control response (Mouchiroud et al., 2013). Separately, C. elegans lifespan can be extended by respiratory chain impairment which also induces mitochondrial protein quality control systems (Durieux et al., 2011). These studies offer strong evidence that the aging process in lower organisms and perhaps the development of age-related disease in mammals is tightly linked to protein quality control, one component of which may be sirtuin-mediated deacylation.
Sirtuins are often implicated in the physiological adaptations to calorie restriction (CR), which extends lifespan and delays the onset of age-related disease in mammals (Guarente, 2013). Interestingly, gene expression profiling of CR mice suggests that cellular mechanisms of proteostasis and macromolecular quality control are key mediators of the CR-induced delay in age-associated pathological changes (Lee et al., 1999). Thus, the involvement of sirtuins in mediating the beneficial effects of CR is consistent with the model that sirtuins support health by removing harmful non-enzymatic acyl-modifications and enhancing protein quality control.
Future Studies
While this model might be useful for understanding the role of acetylation as a post-translational modification compared to other signaling modifications (i.e. phosphorylation), several questions and challenges remain. For example, this model relies on the concept of non-enzymatic acylation, which is challenging to test experimentally. Even in the most rigorous in vivo studies, the possibility of enzymatic transfer by a not-yet-discovered transferase remains. However, despite similar experimental challenges for protein modifications like glycation and carbamylation, these reactions have become accepted as non-enzymatic due to the large body of evidence supporting their reactive nature and their correlation with disease states in which the reactive metabolite is elevated.
In the setting of the cytoplasm or nucleus, where acyltransferases play well-established roles in acetylating several cellular proteins, disentangling the balance between enzymatic and non-enzymatic acyl transfer will prove to be difficult. Based on in vitro studies, the mitochondrial compartment has a pH that is permissive for non-enzymatic acetylation, whereas the cytoplasmic and nuclear compartments are less permissive. However, again based on in vitro evidence, all three compartments could be permissive for non-enzymatic acylation, such as succinylation (Wagner and Payne, 2013). Therefore, the difference between enzymatic and non-enzymatic acylation with respect to sub-cellular compartments as well as the difference in these events on regulation and signaling, will be an important area of future study.
Conclusions
If life originated as a primordial soup of stochastic organic chemical reactions, it is not surprising that some unintended reactions occur between metabolites and proteins. In this setting, the endogenous formation of proteotoxic carbon conjugates is an ongoing and unavoidable consequence of carbon metabolism, which can contribute to protein dysfunction and disease. To prevent and/or reverse this, a variety of cellular quality control and stress resistance programs have evolved. We propose that the sirtuin enzymes are part of the global protein quality control family, and they serve to remove suppressive and potentially damaging non-enzymatic acyl-modifications. This conceptual framework could explain the broad physiological roles of the sirtuins in stress resistance, healthspan, and in some cases organismal lifespan. Continued work in this exciting area will surely uncover new areas of biology regulated by the sirtuin family, and how a myriad of cellular processes are regulated by acylation and deacylation.
Acknowledgements
We would like to acknowledge Chris Newgard, Deb Muoio, and the entire Hirschey Lab for thoughtful feedback. Work in the Hirschey lab is supported by the American Heart Association grants 12SDG8840004 and 12IRG9010008, the Duke O’Brien Center for Kidney Research (5P30DK096493-02), The Edward Mallinckrodt, Jr. Foundation, FARA, The Ellison Medical Foundation, and the National Institutes of Health (grant AA022146 and AG045351).
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- Albaugh BN, Arnold KM, Denu JM. KAT(ching) metabolism by the tail: insight into the links between lysine acetyltransferases and metabolism. Chembiochem : a European journal of chemical biology. 2011;12:290–298. doi: 10.1002/cbic.201000438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Allfrey VG, Faulkner R, Mirsky AE. Acetylation And Methylation Of Histones And Their Possible Role In The Regulation Of RNA Synthesis. Proc Natl Acad Sci U S A. 1964;51:786–794. doi: 10.1073/pnas.51.5.786. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baddiley J, Kekwick RA, Thain EM. A new method for acetylating proteins. Nature. 1952;170:968–970. doi: 10.1038/170968a0. [DOI] [PubMed] [Google Scholar]
- Bharathi SS, Zhang Y, Mohsen AW, Uppala R, Balasubramani M, Schreiber E, Uechi G, Beck ME, Rardin MJ, Vockley J, et al. SIRT3 Regulates Long-chain Acyl-CoA Dehydrogenase by Deacetylating Conserved Lysines Near the Active Site. The Journal of biological chemistry. 2013 doi: 10.1074/jbc.M113.510354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brown K, Xie S, Qiu X, Mohrin M, Shin J, Liu Y, Zhang D, Scadden DT, Chen D. SIRT3 reverses aging-associated degeneration. Cell Rep. 2013;3:319–327. doi: 10.1016/j.celrep.2013.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brownlee M. Biochemistry and molecular cell biology of diabetic complications. Nature. 2001;414:813–820. doi: 10.1038/414813a. [DOI] [PubMed] [Google Scholar]
- Casey JR, Grinstein S, Orlowski J. Sensors and regulators of intracellular pH. Nat Rev Mol Cell Biol. 2010;11:50–61. doi: 10.1038/nrm2820. [DOI] [PubMed] [Google Scholar]
- Cheng HL, Mostoslavsky R, Saito S, Manis JP, Gu Y, Patel P, Bronson R, Appella E, Alt FW, Chua KF. Developmental defects and p53 hyperacetylation in Sir2 homolog (SIRT1)-deficient mice. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:10794–10799. doi: 10.1073/pnas.1934713100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choudhary C, Kumar C, Gnad F, Nielsen M, Rehman M, Walther T, Olsen J, Mann M. Lysine Acetylation Targets Protein Complexes and Co-Regulates Major Cellular Functions. Science. 2009;325:834–840. doi: 10.1126/science.1175371. [DOI] [PubMed] [Google Scholar]
- d’Alayer J, Expert-Bezancon N, Beguin P. Time- and temperature-dependent acetylation of the chemokine RANTES produced in recombinant Escherichia coli. Protein expression and purification. 2007;55:9–16. doi: 10.1016/j.pep.2007.04.022. [DOI] [PubMed] [Google Scholar]
- Dalle-Donne I, Giustarini D, Colombo R, Rossi R, Milzani A. Protein carbonylation in human diseases. Trends Mol Med. 2003;9:169–176. doi: 10.1016/s1471-4914(03)00031-5. [DOI] [PubMed] [Google Scholar]
- Davies KJ. Oxidative stress, antioxidant defenses, and damage removal, repair, and replacement systems. IUBMB Life. 2000;50:279–289. doi: 10.1080/713803728. [DOI] [PubMed] [Google Scholar]
- De Duve C. Vital dust : life as a cosmic imperative. Basic Books; New York: 1995. [Google Scholar]
- De Luca A, Sanna F, Sallese M, Ruggiero C, Grossi M, Sacchetta P, Rossi C, De Laurenzi V, Di Ilio C, Favaloro B. Methionine sulfoxide reductase A down-regulation in human breast cancer cells results in a more aggressive phenotype. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:18628–18633. doi: 10.1073/pnas.1010171107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Delpech M, Levy-Favatier F, Kruh J. In vitro, non enzymatic labelling of histone H1 with [14C] acetyl CoA. Biochimie. 1983;65:291–294. doi: 10.1016/s0300-9084(83)80281-8. [DOI] [PubMed] [Google Scholar]
- Du J, Zhou Y, Su X, Yu JJ, Khan S, Jiang H, Kim J, Woo J, Kim JH, Choi BH, et al. Sirt5 is a NAD-dependent protein lysine demalonylase and desuccinylase. Science. 2011;334:806–809. doi: 10.1126/science.1207861. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Durieux J, Wolff S, Dillin A. The cell-non-autonomous nature of electron transport chain-mediated longevity. Cell. 2011;144:79–91. doi: 10.1016/j.cell.2010.12.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Evans DC, Watt AP, Nicoll-Griffith DA, Baillie TA. Drug-protein adducts: an industry perspective on minimizing the potential for drug bioactivation in drug discovery and development. Chemical Research in Toxicology. 2004;17:3–16. doi: 10.1021/tx034170b. [DOI] [PubMed] [Google Scholar]
- Fan J, Shan C, Kang HB, Elf S, Xie J, Tucker M, Gu TL, Aguiar M, Lonning S, Chen H, et al. Tyr Phosphorylation of PDP1 Toggles Recruitment between ACAT1 and SIRT3 to Regulate the Pyruvate Dehydrogenase Complex. Mol Cell. 2014 doi: 10.1016/j.molcel.2013.12.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feldman JL, Baeza J, Denu JM. Activation of the Protein Deacetylase SIRT6 by Long-chain Fatty Acids and Widespread Deacylation by Mammalian Sirtuins. Journal of Biological Chemistry. 2013 doi: 10.1074/jbc.C113.511261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Finkel T, Deng CX, Mostoslavsky R. Recent progress in the biology and physiology of sirtuins. Nature. 2009;460:587–591. doi: 10.1038/nature08197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Finley LW, Carracedo A, Lee J, Souza A, Egia A, Zhang J, Teruya-Feldstein J, Moreira PI, Cardoso SM, Clish CB, et al. SIRT3 opposes reprogramming of cancer cell metabolism through HIF1alpha destabilization. Cancer Cell. 2011;19:416–428. doi: 10.1016/j.ccr.2011.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fluckiger R, Harmon W, Meier W, Loo S, Gabbay KH. Hemoglobin carbamylation in uremia. New England Journal of Medicine. 1981;304:823–827. doi: 10.1056/NEJM198104023041406. [DOI] [PubMed] [Google Scholar]
- Friedmann DR, Marmorstein R. Structure and mechanism of non-histone protein acetyltransferase enzymes. The FEBS journal. 2013;280:5570–5581. doi: 10.1111/febs.12373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Frye RA. Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochemical and Biophysical Research Communications. 2000;273:793–798. doi: 10.1006/bbrc.2000.3000. [DOI] [PubMed] [Google Scholar]
- Fukami K, Yamagishi SI, Okuda S. Role of AGEs-RAGE system in cardiovascular disease. Curr Pharm Des. 2013 doi: 10.2174/13816128113199990475. [DOI] [PubMed] [Google Scholar]
- Garland PB, Shepherd D, Yates DW. Steady-state concentrations of coenzyme A, acetyl-coenzyme A and long-chain fatty acyl-coenzyme A in rat-liver mitochondria oxidizing palmitate. Biochemical Journal. 1965;97:587–594. doi: 10.1042/bj0970587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garlick RL, Mazer JS, Chylack LT, Jr., Tung WH, Bunn HF. Nonenzymatic glycation of human lens crystallin. Effect of aging and diabetes mellitus. The Journal of clinical investigation. 1984;74:1742–1749. doi: 10.1172/JCI111592. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ghanta S, Grossmann RE, Brenner C. Mitochondrial protein acetylation as a cell-intrinsic, evolutionary driver of fat storage: Chemical and metabolic logic of acetyl-lysine modifications. Crit Rev Biochem Mol Biol. 2013 doi: 10.3109/10409238.2013.838204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guan KL, Xiong Y. Regulation of intermediary metabolism by protein acetylation. Trends Biochem Sci. 2011;36:108–116. doi: 10.1016/j.tibs.2010.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guarente L. Calorie restriction and sirtuins revisited. Genes and Development. 2013;27:2072–2085. doi: 10.1101/gad.227439.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hafner AV, Dai J, Gomes AP, Xiao CY, Palmeira CM, Rosenzweig A, Sinclair DA. Regulation of the mPTP by SIRT3-mediated deacetylation of CypD at lysine 166 suppresses age-related cardiac hypertrophy. Aging. 2010;2:914–923. doi: 10.18632/aging.100252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hall JA, Dominy JE, Lee Y, Puigserver P. The sirtuin family’s role in aging and age-associated pathologies. J Clin Invest. 2013;123:973–979. doi: 10.1172/JCI64094. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hallows WC, Lee S, Denu JM. Sirtuins deacetylate and activate mammalian acetyl-CoA synthetases. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:10230–10235. doi: 10.1073/pnas.0604392103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hansford RG, Johnson RN. The steady state concentrations of coenzyme A SH and coenzyme A thioester, citrate, and isocitrate during tricarboxylate cycle oxidations in rabbit heart mitochondria. Journal of Biological Chemistry. 1975;250:8361–8375. [PubMed] [Google Scholar]
- Hausladen A, Privalle CT, Keng T, DeAngelo J, Stamler JS. Nitrosative stress: activation of the transcription factor OxyR. Cell. 1996;86:719–729. doi: 10.1016/s0092-8674(00)80147-6. [DOI] [PubMed] [Google Scholar]
- Hebert AS, Dittenhafer-Reed KE, Yu W, Bailey DJ, Selen ES, Boersma MD, Carson JJ, Tonelli M, Balloon AJ, Higbee AJ, et al. Calorie restriction and SIRT3 trigger global reprogramming of the mitochondrial protein acetylome. Molecular cell. 2013;49:186–199. doi: 10.1016/j.molcel.2012.10.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirschey M, Aouizerat B, Jing E, Shimazu T, Grueter C, Collins A, Stevens R, Lam M, Muehlbauer M, Schwer B, et al. SIRT3 Deficiency and Mitochondrial Protein Hyperacetylation Accelerate the Development of the Metabolic Syndrome. Molecular Cell. 2011;44:177–190. doi: 10.1016/j.molcel.2011.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirschey MD. Old Enzymes, New Tricks: Sirtuins Are NAD(+)-Dependent De acylases. Cell Metabolism. 2011;14:718–719. doi: 10.1016/j.cmet.2011.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirschey MD, Shimazu T, Goetzman E, Jing E, Schwer B, Lombard DB, Grueter CA, Harris C, Biddinger S, Ilkayeva OR, et al. SIRT3 regulates mitochondrial fatty-acid oxidation by reversible enzyme deacetylation. Nature. 2010;464:121–125. doi: 10.1038/nature08778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Horng H, Benet LZ. The Non-enzymatic Reactivity of the Acyl-linked Metabolites of Mefenamic Acid Towards Amino and Thiol Functional Group Biomolecules. Drug Metab Dispos. 2013 doi: 10.1124/dmd.113.053223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Houtkooper RH, Mouchiroud L, Ryu D, Moullan N, Katsyuba E, Knott G, Williams RW, Auwerx J. Mitonuclear protein imbalance as a conserved longevity mechanism. Nature. 2013;497:451–457. doi: 10.1038/nature12188. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imai S, Armstrong CM, Kaeberlein M, Guarente L. Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature. 2000;403:795–800. doi: 10.1038/35001622. [DOI] [PubMed] [Google Scholar]
- Isenberg I. Histones. Annu Rev Biochem. 1979;48:159–191. doi: 10.1146/annurev.bi.48.070179.001111. [DOI] [PubMed] [Google Scholar]
- Jiang H, Khan S, Wang Y, Charron G, He B, Sebastian C, Du J, Kim R, Ge E, Mostoslavsky R, et al. SIRT6 regulates TNF-alpha secretion through hydrolysis of long-chain fatty acyl lysine. Nature. 2013;496:110–113. doi: 10.1038/nature12038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang W, Wang S, Xiao M, Lin Y, Zhou L, Lei Q, Xiong Y, Guan KL, Zhao S. Acetylation regulates gluconeogenesis by promoting PEPCK1 degradation via recruiting the UBR5 ubiquitin ligase. Mol Cell. 2011;43:33–44. doi: 10.1016/j.molcel.2011.04.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaeberlein M, McVey M, Guarente L. The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev. 1999;13:2570–2580. doi: 10.1101/gad.13.19.2570. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koppaka V, Thompson DC, Chen Y, Ellermann M, Nicolaou KC, Juvonen RO, Petersen D, Deitrich RA, Hurley TD, Vasiliou V. Aldehyde dehydrogenase inhibitors: a comprehensive review of the pharmacology, mechanism of action, substrate specificity, and clinical application. Pharmacol Rev. 2012;64:520–539. doi: 10.1124/pr.111.005538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Langevin F, Crossan GP, Rosado IV, Arends MJ, Patel KJ. Fancd2 counteracts the toxic effects of naturally produced aldehydes in mice. Nature. 2011;475:53–58. doi: 10.1038/nature10192. [DOI] [PubMed] [Google Scholar]
- Lee CK, Klopp RG, Weindruch R, Prolla TA. Gene expression profile of aging and its retardation by caloric restriction. Science. 1999;285:1390–1393. doi: 10.1126/science.285.5432.1390. [DOI] [PubMed] [Google Scholar]
- Li Y, Huang TT, Carlson EJ, Melov S, Ursell PC, Olson JL, Noble LJ, Yoshimura MP, Berger C, Chan PH, et al. Dilated cardiomyopathy and neonatal lethality in mutant mice lacking manganese superoxide dismutase. Nature Genetics. 1995;11:376–381. doi: 10.1038/ng1295-376. [DOI] [PubMed] [Google Scholar]
- Lin R, Tao R, Gao X, Li T, Zhou X, Guan KL, Xiong Y, Lei QY. Acetylation stabilizes ATP-citrate lyase to promote lipid biosynthesis and tumor growth. Mol Cell. 2013;51:506–518. doi: 10.1016/j.molcel.2013.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lowe DM, Tubbs PK. Succinylation and inactivation of 3-hydroxy-3-methylglutaryl-CoA synthase by succinyl-CoA and its possible relevance to the control of ketogenesis. The Biochemical journal. 1985;232:37–42. doi: 10.1042/bj2320037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Madsen AS, Olsen CA. Profiling of substrates for zinc-dependent lysine deacylase enzymes: HDAC3 exhibits decrotonylase activity in vitro. Angewandte Chemie. 2012;51:9083–9087. doi: 10.1002/anie.201203754. [DOI] [PubMed] [Google Scholar]
- Maillard LC. Action of Amino Acids on Sugars. Formation of Melanoidins in a Methodical Way. Compt Rend. 1912;154:66. [Google Scholar]
- Meyer Y, Buchanan BB, Vignols F, Reichheld JP. Thioredoxins and glutaredoxins: unifying elements in redox biology. Annual Review of Genetics. 2009;43:335–367. doi: 10.1146/annurev-genet-102108-134201. [DOI] [PubMed] [Google Scholar]
- Michishita E, Park JY, Burneskis JM, Barrett JC, Horikawa I. Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Mol Biol Cell. 2005;16:4623–4635. doi: 10.1091/mbc.E05-01-0033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moellering RE, Cravatt BF. Functional Lysine Modification by an Intrinsically Reactive Primary Glycolytic Metabolite. Science. 2013;341:549–553. doi: 10.1126/science.1238327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moskovitz J, Bar-Noy S, Williams WM, Requena J, Berlett BS, Stadtman ER. Methionine sulfoxide reductase (MsrA) is a regulator of antioxidant defense and lifespan in mammals. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:12920–12925. doi: 10.1073/pnas.231472998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mouchiroud L, Houtkooper RH, Moullan N, Katsyuba E, Ryu D, Canto C, Mottis A, Jo YS, Viswanathan M, Schoonjans K, et al. The NAD(+)/Sirtuin Pathway Modulates Longevity through Activation of Mitochondrial UPR and FOXO Signaling. Cell. 2013;154:430–441. doi: 10.1016/j.cell.2013.06.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Muller FL, Lustgarten MS, Jang Y, Richardson A, Van Remmen H. Trends in oxidative aging theories. Free Radical Biology and Medicine. 2007;43:477–503. doi: 10.1016/j.freeradbiomed.2007.03.034. [DOI] [PubMed] [Google Scholar]
- Nakagawa T, Lomb DJ, Haigis MC, Guarente L. SIRT5 Deacetylates carbamoyl phosphate synthetase 1 and regulates the urea cycle. Cell. 2009;137:560–570. doi: 10.1016/j.cell.2009.02.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Narayan N, Lee IH, Borenstein R, Sun J, Wong R, Tong G, Fergusson MM, Liu J, Rovira, Cheng HL, et al. The NAD-dependent deacetylase SIRT2 is required for programmed necrosis. Nature. 2012;492:199–204. doi: 10.1038/nature11700. [DOI] [PubMed] [Google Scholar]
- Newman JC, He W, Verdin E. Mitochondrial Protein Acylation and Intermediary Metabolism: Regulation by Sirtuins and Implications for Metabolic Disease. J Biol Chem. 2012 doi: 10.1074/jbc.R112.404863. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paik WK, Pearson D, Lee HW, Kim S. Nonenzymatic acetylation of histones with acetyl-CoA. Biochimica et Biophysica Acta. 1970;213:513–522. doi: 10.1016/0005-2787(70)90058-4. [DOI] [PubMed] [Google Scholar]
- Park J, Chen Y, Tishkoff DX, Peng C, Tan M, Dai L, Xie Z, Zhang Y, Zwaans BM, Skinner ME, et al. SIRT5-Mediated Lysine Desuccinylation Impacts Diverse Metabolic Pathways. Molecular Cell. 2013;50:919–930. doi: 10.1016/j.molcel.2013.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peng C, Lu Z, Xie Z, Cheng Z, Chen Y, Tan M, Luo H, Zhang Y, He W, Yang K, et al. The first identification of lysine malonylation substrates and its regulatory enzyme. Molecular & cellular proteomics : MCP. 2011;10 doi: 10.1074/mcp.M111.012658. M111 012658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Powers ET, Morimoto RI, Dillin A, Kelly JW, Balch WE. Biological and chemical approaches to diseases of proteostasis deficiency. Annual Review of Biochemistry. 2009;78:959–991. doi: 10.1146/annurev.biochem.052308.114844. [DOI] [PubMed] [Google Scholar]
- Quant PA, Tubbs PK, Brand MD. Glucagon activates mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase in vivo by decreasing the extent of succinylation of the enzyme. European journal of biochemistry / FEBS. 1990;187:169–174. doi: 10.1111/j.1432-1033.1990.tb15291.x. [DOI] [PubMed] [Google Scholar]
- Rabbani N, Thornalley PJ. Methylglyoxal, glyoxalase 1 and the dicarbonyl proteome. Amino Acids. 2012;42:1133–1142. doi: 10.1007/s00726-010-0783-0. [DOI] [PubMed] [Google Scholar]
- Rajasekaran NS, Connell P, Christians ES, Yan LJ, Taylor RP, Orosz A, Zhang XQ, Stevenson TJ, Peshock RM, Leopold JA, et al. Human alpha B-crystallin mutation causes oxido-reductive stress and protein aggregation cardiomyopathy in mice. Cell. 2007;130:427–439. doi: 10.1016/j.cell.2007.06.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rardin MJ, He W, Nishida Y, Newman JC, Carrico C, Danielson SR, Guo A, Gut P, Sahu AK, Li B, et al. SIRT5 regulates the mitochondrial lysine succinylome and metabolic networks. Cell Metabolism. 2013a;18:920–933. doi: 10.1016/j.cmet.2013.11.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rardin MJ, Newman JC, Held JM, Cusack MP, Sorensen DJ, Li B, Schilling B, Mooney SD, Kahn CR, Verdin E, et al. Label-free quantitative proteomics of the lysine acetylome in mitochondria identifies substrates of SIRT3 in metabolic pathways. Proc Natl Acad Sci U S A. 2013b;110:6601–6606. doi: 10.1073/pnas.1302961110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rine J, Strathern JN, Hicks JB, Herskowitz I. A suppressor of mating-type locus mutations in Saccharomyces cerevisiae: evidence for and identification of cryptic mating-type loci. Genetics. 1979;93:877–901. doi: 10.1093/genetics/93.4.877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sallustio BC, Nunthasomboon S, Drogemuller CJ, Knights KM. In vitro covalent binding of nafenopin-CoA to human liver proteins. Toxicol Appl Pharmacol. 2000;163:176–182. doi: 10.1006/taap.1999.8868. [DOI] [PubMed] [Google Scholar]
- Scott I, Webster BR, Li JH, Sack MN. Identification of a molecular component of the mitochondrial acetyltransferase programme: a novel role for GCN5L1. The Biochemical journal. 2012;443:655–661. doi: 10.1042/BJ20120118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seiler A, Schneider M, Forster H, Roth S, Wirth EK, Culmsee C, Plesnila N, Kremmer E, Radmark O, Wurst W, et al. Glutathione peroxidase 4 senses and translates oxidative stress into 12/15-lipoxygenase dependent- and AIF-mediated cell death. Cell Metabolism. 2008;8:237–248. doi: 10.1016/j.cmet.2008.07.005. [DOI] [PubMed] [Google Scholar]
- Shimazu T, Hirschey MD, Hua L, Dittenhafer-Reed KE, Schwer B, Lombard DB, Li Y, Bunkenborg J, Alt FW, Denu JM, et al. SIRT3 deacetylates mitochondrial 3-hydroxy-3-methylglutaryl CoA synthase 2 and regulates ketone body production. Cell Metabolism. 2010;12:654–661. doi: 10.1016/j.cmet.2010.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Someya S, Yu W, Hallows WC, Xu J, Vann JM, Leeuwenburgh C, Tanokura M, Denu JM, Prolla TA. Sirt3 mediates reduction of oxidative damage and prevention of age-related hearing loss under caloric restriction. Cell. 2010;143:802–812. doi: 10.1016/j.cell.2010.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stark GR. Reactions of cyanate with functional groups of proteins. 3. Reactions with amino and carboxyl groups. Biochemistry. 1965;4:1030–1036. doi: 10.1021/bi00882a008. [DOI] [PubMed] [Google Scholar]
- Strahl BD, Allis CD. The language of covalent histone modifications. Nature. 2000;403:41–45. doi: 10.1038/47412. [DOI] [PubMed] [Google Scholar]
- Tan M, Peng C, Anderson KA, Chhoy P, Xie Z, Dai L, Park JS, Chen Y, Huang H. Lysine Glutarylation is a Protein Post-translational Modification Regulated By SIRT5. Cell Metab. 2014 doi: 10.1016/j.cmet.2014.03.014. In Press. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tanner KG, Langer MR, Kim Y, Denu JM. Kinetic mechanism of the histone acetyltransferase GCN5 from yeast. Journal of Biological Chemistry. 2000;275:22048–22055. doi: 10.1074/jbc.M002893200. [DOI] [PubMed] [Google Scholar]
- Tanner KG, Trievel RC, Kuo MH, Howard RM, Berger SL, Allis CD, Marmorstein R, Denu JM. Catalytic mechanism and function of invariant glutamic acid 173 from the histone acetyltransferase GCN5 transcriptional coactivator. Journal of Biological Chemistry. 1999;274:18157–18160. doi: 10.1074/jbc.274.26.18157. [DOI] [PubMed] [Google Scholar]
- Tao R, Coleman MC, Pennington JD, Ozden O, Park SH, Jiang H, Kim HS, Flynn CR, Hill S, Hayes McDonald W, et al. Sirt3-mediated deacetylation of evolutionarily conserved lysine 122 regulates MnSOD activity in response to stress. Mol Cell. 2010;40:893–904. doi: 10.1016/j.molcel.2010.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thornalley PJ. Dicarbonyl intermediates in the maillard reaction. Annals of the New York Academy of Sciences. 2005;1043:111–117. doi: 10.1196/annals.1333.014. [DOI] [PubMed] [Google Scholar]
- Toiber D, Erdel F, Bouazoune K, Silberman DM, Zhong L, Mulligan P, Sebastian C, Cosentino C, Martinez-Pastor B, Giacosa S, et al. SIRT6 recruits SNF2H to DNA break sites, preventing genomic instability through chromatin remodeling. Molecular Cell. 2013;51:454–468. doi: 10.1016/j.molcel.2013.06.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ulrich P, Cerami A. Protein glycation, diabetes, and aging. Recent Prog Horm Res. 2001;56:1–21. doi: 10.1210/rp.56.1.1. [DOI] [PubMed] [Google Scholar]
- Verma M, Shulga N, Pastorino JG. Sirtuin-3 Modulates Bak/Bax Dependent Apoptosis. J Cell Sci. 2012 doi: 10.1242/jcs.115188. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- Wagner GR, Payne RM. Widespread and enzyme-independent N{epsilon}-acetylation and N{epsilon}-succinylation in the chemical conditions of the mitochondrial matrix. Journal of Biological Chemistry. 2013 doi: 10.1074/jbc.M113.486753. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Z, Nicholls SJ, Rodriguez ER, Kummu O, Horkko S, Barnard J, Reynolds WF, Topol EJ, DiDonato JA, Hazen SL. Protein carbamylation links inflammation, smoking, uremia and atherogenesis. Nature Medicine. 2007;13:1176–1184. doi: 10.1038/nm1637. [DOI] [PubMed] [Google Scholar]
- Weinert BT, Iesmantavicius V, Moustafa T, Scholz C, Wagner SA, Magnes C, Zechner R, Choudhary C. Acetylation dynamics and stoichiometry in Saccharomyces cerevisiae. Mol Syst Biol. 2013a;10:716. doi: 10.1002/msb.134766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weinert BT, Iesmantavicius V, Wagner SA, Scholz C, Gummesson B, Beli P, Nystrom T, Choudhary C. Acetyl-phosphate is a critical determinant of lysine acetylation in E. coli. Molecular Cell. 2013b;51:265–272. doi: 10.1016/j.molcel.2013.06.003. [DOI] [PubMed] [Google Scholar]
- Weinert BT, Scholz C, Wagner SA, Iesmantavicius V, Su D, Daniel JA, Choudhary C. Lysine succinylation is a frequently occurring modification in prokaryotes and eukaryotes and extensively overlaps with acetylation. Cell Rep. 2013c;4:842–851. doi: 10.1016/j.celrep.2013.07.024. [DOI] [PubMed] [Google Scholar]
- Westerheide SD, Anckar J, Stevens SM, Jr., Sistonen L, Morimoto RI. Stress-inducible regulation of heat shock factor 1 by the deacetylase SIRT1. Science. 2009;323:1063–1066. doi: 10.1126/science.1165946. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu R, Haas W, Dephoure N, Huttlin EL, Zhai B, Sowa ME, Gygi SP. A large-scale method to measure absolute protein phosphorylation stoichiometries. Nat Methods. 2011;8:677–683. doi: 10.1038/nmeth.1636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yi CH, Pan H, Seebacher J, Jang IH, Hyberts SG, Heffron GJ, Vander Heiden MG, Yang R, Li F, Locasale JW, et al. Metabolic regulation of protein N-alpha-acetylation by Bcl-xL promotes cell survival. Cell. 2011;146:607–620. doi: 10.1016/j.cell.2011.06.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yokoyama A, Muramatsu T, Ohmori T, Yokoyama T, Okuyama K, Takahashi H, Hasegawa Y, Higuchi S, Maruyama K, Shirakura K, et al. Alcohol-related cancers and aldehyde dehydrogenase-2 in Japanese alcoholics. Carcinogenesis. 1998;19:1383–1387. doi: 10.1093/carcin/19.8.1383. [DOI] [PubMed] [Google Scholar]
- Yu J, Sadhukhan S, Noriega LG, Moullan N, He B, Weiss RS, Lin H, Schoonjans K, Auwerx J. Metabolic Characterization of a Sirt5 deficient mouse model. Scientific reports. 2013;3:2806. doi: 10.1038/srep02806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu W, Dittenhafer-Reed KE, Denu JM. SIRT3 Protein Deacetylates Isocitrate Dehydrogenase 2 (IDH2) and Regulates Mitochondrial Redox Status. J Biol Chem. 2012;287:14078–14086. doi: 10.1074/jbc.M112.355206. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yuan H, Su L, Chen WY. The emerging and diverse roles of sirtuins in cancer: a clinical perspective. OncoTargets and therapy. 2013;6:1399–1416. doi: 10.2147/OTT.S37750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Z, Tan M, Xie Z, Dai L, Chen Y, Zhao Y. Identification of lysine succinylation as a new post-translational modification. Nature chemical biology. 2010;7:58–63. doi: 10.1038/nchembio.495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao D, Zou SW, Liu Y, Zhou X, Mo Y, Wang P, Xu YH, Dong B, Xiong Y, Lei QY, et al. Lysine-5 acetylation negatively regulates lactate dehydrogenase A and is decreased in pancreatic cancer. Cancer Cell. 2013;23:464–476. doi: 10.1016/j.ccr.2013.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]