Next Article in Journal
Cardiac ECM: Its Epigenetic Regulation and Role in Heart Development and Repair
Next Article in Special Issue
Editorial of Special Issue “Sirtuins in Health and Disease”
Previous Article in Journal
Injectable Gel Form of a Decellularized Bladder Induces Adipose-Derived Stem Cell Differentiation into Smooth Muscle Cells In Vitro
Previous Article in Special Issue
The Role of Sirtuins in Kidney Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Molecular Perspective on Sirtuin Activity

by
Carla S. S. Teixeira
1,
Nuno M. F. S. A. Cerqueira
1,
Pedro Gomes
2,3,4,5 and
Sérgio F. Sousa
1,*
1
UCIBIO/REQUIMTE, BioSIM - Department of Biomedicine, Faculty of Medicine, University of Porto, Alameda Prof. Hernâni Monteiro, 4200-319 Porto, Portugal
2
Department of Biomedicine, Faculty of Medicine, University of Porto, Alameda Prof. Hernâni Monteiro, 4200-319 Porto, Portugal
3
Center for Health Technology and Services Research (CINTESIS), University of Porto, R. Dr. Plácido da Costa, 4200-450 Porto, Portugal
4
Institute of Pharmacology and Experimental Therapeutics, Coimbra Institute for Clinical and Biomedical Research (iCBR), Faculty of Medicine, University of Coimbra, Azinhaga Santa Comba, Celas, 3000-548 Coimbra, Portugal
5
Center for Innovative Biomedicine and Biotechnology (CIBB), University of Coimbra, Azinhaga Santa Comba, Celas, 3000-548 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(22), 8609; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21228609
Submission received: 9 October 2020 / Revised: 7 November 2020 / Accepted: 13 November 2020 / Published: 15 November 2020
(This article belongs to the Special Issue Sirtuins in Health and Disease)

Abstract

:
The protein acetylation of either the α-amino groups of amino-terminal residues or of internal lysine or cysteine residues is one of the major posttranslational protein modifications that occur in the cell with repercussions at the protein as well as at the metabolome level. The lysine acetylation status is determined by the opposing activities of lysine acetyltransferases (KATs) and lysine deacetylases (KDACs), which add and remove acetyl groups from proteins, respectively. A special group of KDACs, named sirtuins, that require NAD+ as a substrate have received particular attention in recent years. They play critical roles in metabolism, and their abnormal activity has been implicated in several diseases. Conversely, the modulation of their activity has been associated with protection from age-related cardiovascular and metabolic diseases and with increased longevity. The benefits of either activating or inhibiting these enzymes have turned sirtuins into attractive therapeutic targets, and considerable effort has been directed toward developing specific sirtuin modulators. This review summarizes the protein acylation/deacylation processes with a special focus on the current developments in the sirtuin research field.

1. Introduction

Proteins are the structural and functional base of all living organisms. To date, the number of proteins that comprise the human proteome is still elusive. The analysis of the human genome shows the existence of approximately 25,000 protein-coding genes [1]. This would suggest the existence of the same number of proteins. Interestingly, to date, more than 90,000 different human proteins have been identified. This discrepancy has been attributed to three district mechanisms: the alternative splicing of precursor mRNAs, single amino acid polymorphisms (SAPs) and posttranslational modifications (PTMs) [2,3,4]. Together, these modifications raise the complexity of the proteomes by two to three orders of magnitude and help to explain the discrepancy between the complexity of vertebrate organisms and the sizes of their encoded genomes [2,5].
There are two general categories of protein PTM: the covalent cleavage of specific peptide bonds from a protein backbone, which can occur by autocatalytic cleavage or can be catalyzed by a group of enzymes named proteases, and the covalent attachment of a chemical group, which is usually an electrophilic fragment of a co-substrate, to a nucleophilic side-chain residue from the protein. The last reaction can be reversible or irreversible and can be enzymatically or non-enzymatically catalyzed.
It has been estimated that about 5% of the genomes of higher eukaryotes are dedicated to enzymes involved in the posttranslational modification of proteins [2]. Currently, there is a record of more than 300 different PTMs [2,6], and the most common include phosphorylation, ubiquitination, alkylation, glycosylation, oxidation and acylation [5,7,8,9,10,11].
The variability that these PTMs bring to proteins is further amplified by the existence of complex regulatory networks involving both positive and negative crosstalk between the different PTMs. This complex PTM crosstalk is the basis of a protein modification code and has significant importance in the regulation of cellular functions [5,12,13,14].
The most extensively studied PTMs are phosphorylation and acylation [9,11]. Among the array of different possible acylation reactions that may occur, acetylation is the most common (for example, more than 20% of mitochondrial proteins are acetylated [15]) [11,13,16,17,18].
The function of protein acylation reactions depends on the acyl group that is attached. It was proposed that the small acyl groups (formyl and acetyl) may function as recognition elements for protein–protein interactions, while long chains of fatty acids can target proteins to membranes and affect signal transduction [19].
The aim of this review is to provide a global overview of the mechanisms underlying both the acylation (Figure 1) and deacylation (Figure 2) of proteins, focusing on a special class of NAD+-dependent protein deacylases named sirtuins.

2. Protein Acylation

Protein acylation consists of the covalent attachment of acyl groups (acetyl, propionyl, butyryl, malonyl, succinyl, crotonyl, 2-hydroxyisobutyryl, glutaryl, benzoyl, myristate, palmitate, farnesyl, geranylgeranyl, etc.) to amino acid residues bearing nucleophilic side chains, such as lysine (-NH2), cysteine (-SH), serine and threonine (-OH), with a specific location in the target protein [19].
The acyl group is generally attached to the side chain of an internal amino acid residue, but acylation can also occur in the protein N-terminus.

2.1. N-Terminal Acylation

The known examples of N-terminal acylation are the N-myristoylation of glycine, the N-palmitoylation of cysteine, and N-terminal acetylation [19]. Among those, the most common and best characterized is N-terminal acetylation [20].

N-Terminal Acetylation

In eukaryotes, the acetylation of an α-amino group of the N-terminus is an irreversible covalent modification catalyzed by N-terminal-acetyltransferases (NATs) that occurs co-translationally in more than 80% of all proteins [21,22]. N-terminal acetylation neutralizes the positive charge of the free amino group and blocks the N-terminus from further modifications. This modification influences several protein characteristics such as folding, lifetime and degradation, subcellular localization, interactions and complex formation [23,24]. Eukaryotes possess different NATs with different substrate specificities (NatA, NatB, NatC, NatD, NatE and NatF) [25]. Depending on the NAT isoform, the acetyl group may be transferred to the non-cleaved initial methionine residue of the nascent polypeptide chain, or the reaction may occur after the excision of the methionine residue by the ribosomally bound methionine aminopeptidases, and the acetyl group is added to the residue that is positioned immediately after the excised methionine.

2.2. Acylation of Internal Lysine Residues

The acylation of internal lysine residues is a reversible and highly regulated PTM that targets many proteins from different cellular compartments, such as the mitochondria, cytosol, nucleus and lumen of the endoplasmic reticulum (ER) [26,27,28]. The target proteins are involved in distinct cellular processes, such as the cell cycle, nuclear transport, chromatin remodeling, mRNA splicing, actin nucleation, signal transduction [26], protein homeostasis and autophagy [27,28].
Although protein acylation can be an activating or inhibitory modification, recent metabolic studies have suggested that, in mitochondria, acylation has a consistently inhibitory role [29,30]
Protein acylation may occur by two distinct mechanisms: it may be catalyzed by acyl-specific transferases that transfer acyl groups from acyl-coenzyme A (acyl-CoA) to the amino acid residues [24,31,32,33], or it may occur non-enzymatically due to the intrinsic reactivity of acyl-CoA thioesters [29,33,34,35].
The first evidence of acylation as a PTM came in 1964 when lysine acetylation was discovered as a PTM of histones [36]. The evidence for the addition of other acyl groups as a PTM was only recently discovered, and therefore, the function and mechanisms behind them are still elusive [7].

Acetylation of Internal Lysine Residues

Similar to terminal acetylation, the addition of an acetyl group to an internal lysine residue increases the size of its side chain and neutralizes the positive charge of its amino group. This modification will raise the protein hydrophobicity and consequently induce significant conformational changes that will ultimately affect several protein properties, including transcriptional activity, DNA–protein interactions, subcellular localization, enzymatic activity, folding, peptide–receptor recognition and protein stability [15].
Enzymatic Nε-lysine acetylation has been deeply studied and requires an acetyl group donor (acetyl-CoA), an acetyl group acceptor (the ε-amino group of an internal lysine residue from a polypeptide chain) and an acetyl-CoA:lysine acetyltransferase (lysine acetyltransferase (KAT)) to catalyze the acetyl exchange. The mechanism behind the enzymatic addition of other acyl groups to Nε-lysine requires further studies [29].
Non-enzymatic protein acetylation was identified decades ago [37], and its occurrence was found to be especially favorable in the mitochondria [38,39,40,41,42].
The particular biochemical properties of the mitochondria, namely, its pH and metabolite concentrations, are proposed to favor the occurrence of other non-enzymatic acylation reactions, namely, succinylation, malonylation and glutarylation [29].

2.3. S-Palmitoylation of Cysteine Residues

The reversible attachment of palmitic acid to an internal cysteine residue (S-palmitoylation) is the most common type of protein fatty acylation in eukaryotic cells [43,44,45]. The attachment of the palmitic acid to the protein is catalyzed by palmitoyl transferases [46,47], and its removal is mediated by palmitoyl thioesterases.
The palmitoyl transferases belong to the zinc-finger DHHC domain-containing protein family that is characterized by the presence of a conserved aspartate–histidine–histidine–cysteine (DHHC) cysteine-rich domain [46,47]. They are integral membrane proteins, with four to six transmembrane domains and the N and C termini present in the cytosol [48], a characteristic that has made difficult their purification and consequently their identification.
In spite of the recent advances in the field, there are still many questions regarding the functioning of DHHC enzymes. To date, their selectivity, their forms of substate selection and the identification of palmitoylation sites are still elusive [44].

2.4. The Influence of Protein Acylation in Human Health and Disease

The attachment of acyl groups to positively charged lysine residues neutralizes the positive charge and consequently affects the protein’s physicochemical properties, ultimately affecting its stability, function, catalytic activity, protein–protein and protein–DNA interactions, degradation and subcellular location [43,44,49].
Many of these post-translationally acylated proteins have relevant roles in vital physiological processes including gene transcription, cell division, cytoskeleton organization, DNA damage repair, DNA replication, signal transduction, protein folding, autophagy, apoptosis, lipid storage and breakdown, mitochondrial fission and fusion, protein synthesis, ion transport, redox and metabolism regulation. [49] Additionally, protein acylation has also been related to the protein aggregation process, which is implicated in several neurological pathologies such as amyotrophic lateral sclerosis [50] and Alzheimer’s disease [51].
Therefore, the correct functioning and regulation of the enzymes involved in protein acylation (and deacylation) are essential for human health, and their misregulation has been associated with several pathologies.
There is evidence of germline mutations in several KATs (for example, KAT6A, KAT6B, CREB- binding protein (CBP) and EP300) that result in disorders, for example, Say–Barber–Biesecker–Young–Simpson (or Ohdo) syndrome [52], Genitopatellar syndrome [53,54] or Rubinstein–Taybi syndrome. These and other disorders caused by KAT mutations are associated with intellectual impairment, developmental delays and physical abnormalities such as facial dysmorphisms [52,53,54,55,56].
Protein acetylation has been associated with several cancers [57], with inflammation and immunity (Falkenberg, 2014, pp. 673–691) and with metabolic diseases such as diabetes [58].
In spite of its evident therapeutic potential, the inhibition of KATs has not been widely explored. A recent study described the development of A-485, a potent and selective catalytic inhibitor of p300 and CREB-binding protein (CBP) that competes with the substrate acetyl-CoA. The compound selectively inhibited cell proliferation in lineage-specific tumor types, including several hematological malignancies and androgen receptor-positive prostate cancer [59].
Another study reported the development of two highly potent and selective inhibitors of KAT6A and KAT6B named WM-8014 and WM-1119, respectively, that are reversible competitors of acetyl-CoA. The inhibition of KAT6A and KAT6B inhibited the growth of lymphoma in mice [60].
Detailed knowledge about the catalytic activity of the different KATs would be important for identifying potent and selective inhibitors for other KATs that would enable obtaining deep knowledge about their biological functions and exploring their therapeutic potential.
Reversible S-palmitoylation is a dynamic process that has several important functions in subcellular protein trafficking, in protein stability (by the prevention of ubiquitination and subsequent degradation) and in the modulation of protein interactions (adhesion and signaling), but its most studied function is its capability to increase the affinity of soluble proteins for lipophilic membranes [61].
The alteration of DHHC expression and the consequent palmitoylation impairment have been associated with several cancers (for example, leukemia, colorectal, hepatocellular and non-small-cell lung cancers) [44,62,63] and with several other disease states resulting from organ-specific processes. One of the most studied cases is the relation between palmitoylation and neuronal functions. There are many reports relating defects in palmitoylation regulation or in the enzymes responsible for palmitoylation and depalmitoylation processes with several neurological disorders such as Alzheimer’s, Parkinson’s or Huntington’s disease, schizophrenia and intellectual disability [44,64].

Protein Acylation and “Carbon Stress”

It has been recently suggested that an increase in the concentration of reactive carbon metabolites, resulting from physiological or pathological situations, can culminate in the abnormal occurrence of non-enzymatic protein acylation reactions [29]. This situation can have a detrimental effect on protein function and ultimately disrupt cellular homeostasis. In these scenarios, the non-enzymatic protein acylations are considered a form of “carbon stress” [29].
Recently, it has been found that under carbon stress situations that lead to an increase in protein acylation, a particular group of deacylases, named sirtuins (Sirts), are called to intervene, and, in some conditions, their expression can be upregulated. The sirtuins are part of a protein quality control response that is vital for reducing protein acylation, ensuring protein quality control and consequently reducing carbon stress. An increase in carbon stress and/or the impairment of the carbon stress response will reduce protein quality control and ultimately reduce protein function, leading to several age-related diseases such as neurodegeneration, diabetes, cardiovascular disease or cancer [18,65,66,67,68,69,70].

3. Protein Deacylation

Efficient and correct deacetylation activity is of major importance for the correct functioning of several vital biological processes, including DNA recombination [71] and repair [72], transcriptional silencing [73], axonal protection [74], fat mobilization [75], apoptosis [76,77] and aging [78,79].
The hydrolysis of acyl groups from acyl-lysine residues is catalyzed by a group of enzymes named lysine deacetylases (KDACs).
The human genome encodes a total of 18 KDACs that can be grouped into two categories or superfamilies: the Zn2+-dependent KDACs or histone deacetylases (HDACs) and the NAD+-dependent sirtuin deacetylases (Figure 2).
It should be noted that although grouped in the same family, some KDACs have different acyl selectivities [80,81,82,83], and others are thought to have poor or no deacetylase activity [84].

3.1. Zn2+-Dependent KDACs

The Zn2+-dependent KDACs (or HDACs) are usually called the classic KDACs and account for 11 of the 18 KDACs. They share a highly conserved deacetylase domain and, based on their phylogenetic conservation and in-sequence similarities, can be divided into four classes, named classes I, IIa, IIb and IV (Figure 2). They differ in their enzymatic function, structure, expression patterns and subcellular localization. Classes I and IV are nuclear, class IIb is cytoplasmic, and class IIa is primarily nuclear but upon the activation of signaling is exported to the cytoplasm [84]. Some in vitro experiments suggest that vertebrate class IIa KDACs show poor catalytic activity, which may be related with the replacement of a conserved tyrosine by a histidine in the catalytic pocket [85].
These enzymes exist in multiprotein complexes that determine their substrate specificity, and their catalysis usually generates deacetylated lysine and acetate [49,84,86].

3.2. NAD+-Dependent Sirtuin Deacetylases

The remaining seven KDACs are named sirtuins and unlike the classic KDACs, they require the cofactor NAD+ as a co-substrate [87].
The name sirtuin came from the family’s founding member, the silent information regulator 2 (Sir2) from Saccharomyces cerevisiae [88].
They comprise the class III family of KDACs and have seven members (Sirt1–Sirt7). Sirtuins are evolutionarily conserved in all domains of life, and, based on their sequence similarity, they are classified into five classes (I–IV and U) (Figure 2). The seven mammalian sirtuin genes are included in classes I to IV: Sirt1, Sirt2 and Sirt3 are in class I, Sirt4 in class II, Sirt5 in class III, and Sirt6 and Sirt7 in class IV (Figure 2) [89]. Sirt1 has the highest sequence homology to Sir2 in yeast, and it is the most studied sirtuin. The different sirtuins are distinguished by their subcellular localization, acyl lysine substrate specificity, enzymatic activity, and biochemical and metabolic functions. Their impairment is related to different metabolic and health issues [90].
Although deacylation is the main activity catalyzed by the sirtuin enzymes, there is evidence that some sirtuins (Sirt1, 4 and 6) can also ADP-ribosylate protein substrates [91].

3.2.1. Sirtuin Structure

Sirtuin proteins contain a conserved catalytic core domain composed of approximately 275 amino acid residues flanked by N- and C-terminal regions with a variable sequence and length [78,92]. The catalytic core domain shows a high degree of structural superposition among the different sirtuins. It adopts an elongated shape containing a classical open α/β Rossmann-fold structure that is characteristic of NAD+/NADH-binding proteins, and a smaller globular domain composed of two insertions in the Rossmann fold. One of these insertions binds a structural zinc ion that is coordinated with four conserved cysteine residues [93]; the other insertion is a helical module (Figure 3a).
Between the two domains exists a deep cleft where the enzyme active site is located and where both NAD+ and acetyl-lysine substrates bind [90]. It was proposed that when the peptide containing the acylated lysine binds to the enzyme’s cleft, its main chain establishes β sheet-like interactions with two flanking strands. One of those strands is positioned in the Rossmann fold. The other one is located in a loop that contains a highly conserved FGExL motif, and is positioned between the Rossmann fold and the Zn2+-binding module [94] (Figure 3b). The formation of this so-called “β staple” interaction, in which the substrate links the Rossmann fold and the Zn2+-binding module, inserts the acetyl-lysine side chain into a conserved, mainly hydrophobic tunnel and consequently changes the enzyme´s conformation from an open to a closed conformation [93,95]. The closed conformation of the enzyme facilitates the correct binding of NAD+ inside a conserved hydrophobic C pocket that is adjacent to the acyl-lysine-binding tunnel. This binding order is important because the occupation of the acetyl-lysine-binding tunnel seems to restrict the binding conformation of NAD+ and force it to adopt a “productive conformation” in which its adenine ring forms extensive hydrogen bonds and van der Waals interactions with the Rossmann-fold domain, and its nicotinamide ring is inserted into the C pocket (Figure 3c) [95,96].

3.2.2. Substrate Specificity of Sirtuins

A few studies demonstrate that sirtuins show a high level of substrate specificity for certain acetylation sites in specific substrates [98,99,100]. Accordingly, substrate recognition by sirtuins is affected by differences in the sirtuins’ binding clefts, by the subcellular localization and by some particular characteristics of the substrate, namely, the acylated residue, the attached acyl group, the three-dimensional structure of the substrate, the substrate sequence and the in vivo interactions of the substrate.

The Influence of Sirtuin Structure

The sirtuin’s active site is positioned in a cleft composed of the Rossmann-fold domain, the Zn2+-binding domain and the four loops that connect the two domains. This cleft is the region of the enzyme that contains the highest sequence conservation within the different sirtuins [98].
The non-Zn2+-binding module from the small Zn2+-binding domain is the area that shows more variability, either in the primary sequence or in the secondary and tertiary structures among the different Sir2 homologues. This observation suggests that this domain may be involved in the sirtuin’s substrate specificity. Additionally, it has also been observed that the small Zn2+-binding domains of the archaeal and bacterial sirtuins have a similar overall topology, while, in eukaryotic sirtuins, they show a higher secondary structure variability [90]. This difference may reflect the higher number of sirtuins that are expressed in eukaryotes and that are required to distinguish a greater number of substrates [101].
Other regions distanced from the active site may also be involved in the sirtuin’s discrimination between different substrates [90].

The Influence of the Subcellular Localization of Sirtuins

The different cellular compartments have different proteins; therefore, the distinct subcellular localizations of the seven sirtuins play an important role in their substrate specificities.
Sirt1 is localized predominantly in the nucleus but was also found in the cytoplasm [102]. Sirt2 is mainly cytosolic but was also found in the nucleus [103]. Sirt3 is predominantly localized in the mitochondrial matrix [104], but there is evidence that it moves from the nucleus to mitochondria during cellular stress [105]. Sirt4 was only detected in the mitochondrial matrix. Sirt5 is predominately mitochondrial but is also active in the cytosol [106]. Currently, Sirt6 and Sirt7 are thought to be nuclear.

The Influence of the Acylated Residue

The existing data suggest that sirtuins only have deacylation activity on acylated lysine residues [107]. However, since our knowledge of their activity is still very limited, the possibility that they can also deacylate other residues should not be excluded.

The Influence of the Acyl Group

In vitro studies showed that only class I sirtuins (Sirt1, 2 and 3) have robust deacetylase activity, although they could also remove long-chain fatty acyl groups. Sirtuins 4–7 have preferences for longer acyl chains [108,109].
Sirt4 was recently shown to remove methylglutaconyl, hydroxymethyl and methylglutaryl from lysine residues [83].
Sirt5 has robust lysine desuccinylase, demalonylase [80,81] and deglutarylase [82] activities. Sirt6 has efficient lysine depalmitoylase and demyristoylase activity [110]. Sirt7 was shown to act as a lysine desuccinylase [111].
Many of these studies were performed in vitro and, therefore, require in vivo validation. The recent development of new methodologies for exploring protein acylation in vivo may bring a new breath of life to this field [19,112].

The Influence of Protein Structure

Previous studies suggested that sirtuins deacetylate lysine residues located in regions without a defined secondary structure or in loop regions of the substrate proteins [113]. Recent studies demonstrated that sirtuins are also able to deacetylate lysine residues located in structured or rigid regions of a protein [114]. One of the major limitations in this field is the difficulty of obtaining the co-crystallized structures of sirtuins with natively folded proteins. Indeed, most studies use small peptides containing acetylated lysine residues to study the enzyme–substrate interaction, an approach that may preclude some important information or even skew the conclusions. To get a full picture of the enzyme–substrate interaction, it would be necessary to increase the number of studies using a structural and functional approach in a context of site-specifically acetylated full-length and natively folded substrate proteins.

The Influence of Protein Sequence

Structural data show that during catalysis, the side chains of the residues preceding and proceeding the acyl lysine interact with both the Rossmann fold and Zn2+-binding module in a “β staple” interaction, as described in the previous section. In order for this enzyme–substrate interaction to be possible and energetically favorable, the side chains of the substrate residues that interact with the enzyme must be chemically and geometrically compatible with the residues that compose the binding cleft of each sirtuin. Because the binding cleft from each sirtuin presents some specific features that distinguish it, the existence of some sequence similarities among the substrates preferentially catalyzed by a given sirtuin would be expected [98,115]. Although some studies have suggested that some sirtuins have a preference for certain amino acid residues in determined sequence positions [98,99,100,114], the results between different studies are not always convergent in their conclusions. A study that performed an analysis of sequences of biochemically confirmed substrates for Sirt2 or Sirt3 concluded that there was no clear consensus in their sequence [93]. On the other hand, some data suggest that the deacetylase activity of Sirt6 is sequence-dependent [110]. In a recent review, it was suggested that this chemical interaction between the enzyme-binding pocket and the substrate residues flanking the acyl lysine could be more important for substrates whose acyl groups show weaker binding affinities for the sirtuin. For acyl groups that bind more tightly to the enzyme´s catalytic pocket, the chemical contribution of those substrate residues would be less important [93].

The Influence of Protein Interactions

Under physiological conditions, sirtuins can interact with other proteins or with DNA, and those interactions influence the sirtuin’s substrate specificity. For example, Sirt1 and Sirt3 bind so tightly to their substrate proteins p53 and AceCS2, respectively, that they coimmunoprecipitate [116,117]. In other cases, the recruitment of a sirtuin for a given substrate or vice versa may be mediated by the interaction with other proteins (e.g., transcription factors). For example, Sirt6 and Sirt7 bind to certain transcription factors, which recruit them to different chromatin regions, where they catalyze the deacetylation of a specific histone at specific target genes [118,119,120].

3.2.3. Catalytic Mechanism of Sirtuin Deacylation

According to the existing data, sirtuins are catalytically active only when the peptide containing the acylated lysine is correctly positioned inside the binding tunnel, the enzyme is in the closed conformation, and NAD+ has its nicotinamide ring inside the hydrophobic C pocket and the α face of its N-ribose ring exposed to the acetyl lysine carbonyl group [95,96].
Although it is well established that the general sirtuin catalytic mechanism proceeds in two consecutive stages (Figure 4), the details about the chemistry involved in the generation of each reaction intermediate are still not consensual [121,122].
In Stage I occurs the ADP-ribosylation of acetyl lysine, which involves the cleavage of the nicotinamide moiety of NAD+ and the nucleophilic attack of the side chain of the acetylated lysine from the protein substrate to form a positively charged ADP-ribosyl-peptidylimidate (or C1′-O-alkylamidate) intermediate and nicotinamide. It was proposed that this reaction occurs through a highly dissociative and concerted displacement mechanism [121]. This reaction is reversible, so NAD+ can be resynthesized when nicotinamide concentrations are elevated in solution.
Computational and experimental evidence has shown that Stage II starts with the deprotonation of the 2′-OH group by a conserved histidine residue that acts as a general base (Step 1 from Figure 4). This facilitates the intra–molecular nucleophilic attack of the 2′ hydroxyl onto the positively charged iminium carbon and culminates in the formation of a bicyclic intermediate. In the second step, which is the reaction-limiting step, occurs the collapse of the bicyclic intermediate, in the presence of a water molecule, generating a tetrahedral intermediate (Steps 2a and 2b in Figure 4). In the third step, there is a proton transfer from the positively charged histidine to the amino group of the tetrahedral intermediate (Step 3 in Figure 4). In the fourth step occurs the breakdown of the tetrahedral intermediate into the reaction products: the deacylated lysine and the 2′-O-acetyl-ADP- ribose (2′-AADPR) (Step 4 in Figure 4) [122,123], which can be non-enzymatically isomerized to 3′-O- acetyl-ADP-ribose.
The reaction products are a mixture of 2′- and 3′-O-acetyl ADP-ribose, nicotinamide and the deacetylated peptide [92,124,125].

3.2.4. The Influence of Sirtuins in Human Health and Disease

The sirtuins are a family of enzymes that target different proteins, including histones, transcription factors or proteins involved in DNA repair. Their variable subcellular distribution and the variability of the substrates allows them to control several vital molecular pathways that are involved in cell survival, neuronal signaling, energy metabolism, tissue regeneration, DNA repair, inflammation or circadian rhythms [94,103,118,119,126,127,128,129]
Caloric restriction is the only effective way to naturally extend lifespan and eventually health span in several organisms including humans [130].
Several studies suggest that caloric restriction increases the expression levels of sirtuins, with the exception of Sirt4 [131,132,133] This relation between sirtuin activation and increased lifespan has suggested that sirtuins may have a role in the beneficial effects elicited by a caloric restriction diet [134]. This assumption has boosted the search for potent sirtuin-activating compounds (STACs) [135].
The abnormal activity and/or expression of several sirtuins has been correlated with several cancer types. Several studies suggest that Sirt1 can act both as a tumor promoter and as a tumor suppressor [136,137]. It regulates many tumor suppressors and DNA repair genes, and its upregulation was corelated with a higher chance of being resistant to chemotherapy [138].
Sirt2, similarly to Sirt1, also has a regulatory function, and it has been suggested that it can act both as a tumor promoter and as a tumor suppressor [139].
Sirt3 was shown to act as a tumor suppressor by inhibiting glycolysis metabolism through the deacetylation and consequent activation of pyruvate dehydrogenase [140], but it is also possible that it can act as a tumor promoter in some situations.
Sirt4 was correlated with the inhibition of the progression of colorectal cancer, and its underexpression was correlated with a worse prognosis [141].
The overexpression of Sirt5 in non-small-cell lung cancer tissues was found to be a marker of low survival [142].
Sirt6 showed a controversial role in several cancers. A reduction in Sirt6 expression was correlated with tumor progression with a poor clinical outcome. Its overexpression was shown to promote oncogenic activity in solid and in hematologic tumors [143].
The overexpression of Sirt7 has been associated with aggressive cancers and low survival, whereas its depletion has been associated with a less aggressive phenotype [120].
Sirt1 is the most studied sirtuin. Alterations of the level of Sirt1 expression were associated with the outcomes of several metabolic and neurodegenerative diseases, cancer and aging. A reduction in Sirt1 expression has been related to cardiovascular and neurodegenerative diseases such as Alzheimer’s and Parkinson’s, and with some metabolic diseases such as obesity and diabetes [144,145,146,147,148]. It has been proposed that the downregulation of Sirt1 along with the disease progression may result from the concomitant increase in oxidative stress and inflammation [146,149].
Some age-related diseases and endocrine system dysfunctions are associated with an increase in Sirt1 expression, albeit with a decrease in its activity. It has been hypothesized that in these cases, the increase in Sirt1 expression is a way to compensate for the decline in Sirt1 activity [144].
The current knowledge about the relationships between sirtuins and certain health and disease conditions strongly suggests that the development of molecules capable of selectively activating each human sirtuin variant may bring health benefits through the stimulation of its anti-inflammatory, cardio-protective, neuroprotective and anti-tumor activities. On the other hand, the relationship between the overexpression of some sirtuin variants and the proliferation of certain cancer cells and the development of some metabolic disorders also suggests that their selective inhibition would also be beneficial in certain disease conditions. Therefore, several activators of Sirt1 and inhibitors of Sirt1 and Sirt2 have been developed and are actually in clinical trials. This issue has been deeply explored in a recent review [150].

4. Conclusions

Although it is now evident that protein acylation is a complex PTM that embraces the addition of a wide range of different acyl groups to specific protein residues, both the acylation and the deacylation processes require deeper investigation. The major limitation of previous studies is the lack of efficient methodologies that are capable of providing an unbiased identification of the acylated proteins, their acylation modification sites and the specific acyl modifications in vivo. As a significant portion of the current knowledge about these mechanisms arose from in vitro studies, it is important to validate those results in vivo both to prove their biological occurrence and to correlate those modifications with their function in the cell and ultimately in the organism.
The correct functioning and regulation of the enzymes that catalyze the addition and the removal of the acyl groups are of major importance for a variety of metabolic processes, and their impairment has already been related to several human pathologies, including neurodegeneration [64,151,152], cancer [62,63,153,154] and cardiovascular diseases [155,156].
The seminal discovery that the upregulation of the Sir2 gene was able to increase the replicative lifespan of yeast [79,157] has sparked great interest in sirtuin biology. From there on, several studies have shown that sirtuins play critical roles in epigenetics, cell death and lifespan regulation [58,127,128,158] and that their abnormal activity is implicated in several diseases, such as cancer, neurodegenerative disorders, obesity and diabetes [129,159].
Recently, it was found that increasing their activity was associated with the delay of some age-related cardiometabolic diseases [160] and could even increase longevity [161,162,163,164]. These findings have turned sirtuins into attractive therapeutic targets, and considerable effort has been directed toward developing specific sirtuin activators and inhibitors. A deep knowledge of sirtuin enzymatic activity and allosteric regulation is imperative for the development of highly specific mechanism-based sirtuin modulators.
The determination of the catalytic mechanisms of acylation and deacylation by each enzyme, and the identification of the specificities that distinguish them among the members of the same families would be of great importance for improving the current knowledge about these enzymes and the associated enzymatic processes. The knowledge of the transition state structures of the rate-limiting steps of each reaction, with atomistic detail, would provide a promising approach for the design of potent and specific molecules with activating or inhibitory characteristics.

Author Contributions

Conceptualization, C.S.S.T. and S.F.S.; investigation, C.S.S.T.; writing—original draft preparation, C.S.S.T.; writing—review and editing, N.M.F.S.A.C., P.G. and S.F.S.; supervision, N.M.F.S.A.C. and S.F.S.; project administration, S.F.S.; funding acquisition, S.F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fundação para a Ciência e a Tecnologia, grants UIDB/04378/2020 and SFRH/BD/114886/2016.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. International Human Genome Sequencing Consortium. Finishing the euchromatic sequence of the human genome. Nature 2004, 431, 931–945. [Google Scholar] [CrossRef]
  2. Walsh, C.T.; Garneau-Tsodikova, S.; Gatto, G.J., Jr. Protein posttranslational modifications: The chemistry of proteome diversifications. Angew. Chem. Int. Ed. Engl. 2005, 44, 7342–7372. [Google Scholar] [CrossRef]
  3. Gilbert, W. Why genes in pieces? Nature 1978, 271, 501. [Google Scholar] [CrossRef]
  4. Roth, M.J.; Forbes, A.J.; Boyne, M.T., 2nd; Kim, Y.B.; Robinson, D.E.; Kelleher, N.L. Precise and parallel characterization of coding polymorphisms, alternative splicing, and modifications in human proteins by mass spectrometry. Mol. Cell. Proteom. 2005, 4, 1002–1008. [Google Scholar] [CrossRef] [Green Version]
  5. Hunter, T. The age of crosstalk: Phosphorylation, ubiquitination, and beyond. Mol. Cell 2007, 28, 730–738. [Google Scholar] [CrossRef]
  6. Cantin, G.T.; Yates, J.R., 3rd. Strategies for shotgun identification of post-translational modifications by mass spectrometry. J. Chromatogr. A 2004, 1053, 7–14. [Google Scholar] [CrossRef]
  7. Nahomi, R.B.; Nandi, S.K.; Rakete, S.; Michel, C.; Fritz, K.S.; Nagaraj, R.H. Lysine malonylation and propionylation are prevalent in human lens proteins. Exp. Eye Res. 2020, 190, 107864. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, Z.; Tan, M.; Xie, Z.; Dai, L.; Chen, Y.; Zhao, Y. Identification of lysine succinylation as a new post-translational modification. Nat. Chem. Biol. 2011, 7, 58–63. [Google Scholar] [CrossRef] [PubMed]
  9. Qausain, S.; Srinivasan, H.; Jamal, S.; Nasiruddin, M.; Khan, M.K.A. Phosphorylation and Acetylation of Proteins as Posttranslational Modification: Implications in Human Health and Associated Diseases. In Protein Modificomics; Academic Press: Cambridge, MA, USA, 2019; pp. 69–86. [Google Scholar] [CrossRef]
  10. Brooks, C.L.; Gu, W. Ubiquitination, phosphorylation and acetylation: The molecular basis for p53 regulation. Curr. Opin. Cell Biol. 2003, 15, 164–171. [Google Scholar] [CrossRef]
  11. Kouzarides, T. Acetylation: A regulatory modification to rival phosphorylation? EMBO J. 2000, 19, 1176–1179. [Google Scholar] [CrossRef]
  12. Caron, C.; Boyault, C.; Khochbin, S. Regulatory cross-talk between lysine acetylation and ubiquitination: Role in the control of protein stability. Bioessays 2005, 27, 408–415. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, X.-J.; Seto, E. Lysine acetylation: Codified crosstalk with other posttranslational modifications. Mol. Cell 2008, 31, 449–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Lu, Z.; Cheng, Z.; Zhao, Y.; Volchenboum, S.L. Bioinformatic analysis and post-translational modification crosstalk prediction of lysine acetylation. PLoS ONE 2011, 6, e28228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kim, S.C.; Sprung, R.; Chen, Y.; Xu, Y.; Ball, H.; Pei, J.; Cheng, T.; Kho, Y.; Xiao, H.; Xiao, L.; et al. Substrate and functional diversity of lysine acetylation revealed by a proteomics survey. Mol. Cell 2006, 23, 607–618. [Google Scholar] [CrossRef]
  16. Lu, J.Y.; Lin, Y.Y.; Sheu, J.C.; Wu, J.T.; Lee, F.J.; Chen, Y.; Lin, M.I.; Chiang, F.T.; Tai, T.Y.; Berger, S.L.; et al. Acetylation of yeast AMPK controls intrinsic aging independently of caloric restriction. Cell 2011, 146, 969–979. [Google Scholar] [CrossRef] [Green Version]
  17. Chuang, C.; Lin, S.H.; Huang, F.; Pan, J.; Josic, D.; Yu-Lee, L.Y. Acetylation of RNA processing proteins and cell cycle proteins in mitosis. J. Proteome Res. 2010, 9, 4554–4564. [Google Scholar] [CrossRef] [Green Version]
  18. Wang, Q.; Zhang, Y.; Yang, C.; Xiong, H.; Lin, Y.; Yao, J.; Li, H.; Xie, L.; Zhao, W.; Yao, Y.; et al. Acetylation of metabolic enzymes coordinates carbon source utilization and metabolic flux. Science 2010, 327, 1004–1007. [Google Scholar] [CrossRef] [Green Version]
  19. Thinon, E.; Hang, H.C. Chemical reporters for exploring protein acylation. Biochem. Soc. Trans. 2015, 43, 253–261. [Google Scholar] [CrossRef] [Green Version]
  20. Deng, S.; Marmorstein, R. Protein N-Terminal Acetylation: Structural Basis, Mechanism, Versatility, and Regulation. Trends Biochem. Sci. 2020. [Google Scholar] [CrossRef]
  21. Brown, J.L.; Roberts, W.K. Evidence that approximately eighty per cent of the soluble proteins from Ehrlich ascites cells are Nalpha-acetylated. J. Biol. Chem. 1976, 251, 1009–1014. [Google Scholar]
  22. Arnesen, T.; Van Damme, P.; Polevoda, B.; Helsens, K.; Evjenth, R.; Colaert, N.; Varhaug, J.E.; Vandekerckhove, J.; Lillehaug, J.R.; Sherman, F.; et al. Proteomics analyses reveal the evolutionary conservation and divergence of N-terminal acetyltransferases from yeast and humans. Proc. Natl. Acad. Sci. USA 2009, 106, 8157–8162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Aksnes, H.; Hole, K.; Arnesen, T. Molecular, cellular, and physiological significance of N-terminal acetylation. Int. Rev. Cell Mol. Biol. 2015, 316, 267–305. [Google Scholar] [CrossRef] [PubMed]
  24. Drazic, A.; Myklebust, L.M.; Ree, R.; Arnesen, T. The world of protein acetylation. Biochim. Biophys. Acta 2016, 1864, 1372–1401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Starheim, K.K.; Gevaert, K.; Arnesen, T. Protein N-terminal acetyltransferases: When the start matters. Trends Biochem. Sci. 2012, 37, 152–161. [Google Scholar] [CrossRef] [PubMed]
  26. Choudhary, C.; Kumar, C.; Gnad, F.; Nielsen, M.L.; Rehman, M.; Walther, T.C.; Olsen, J.V.; Mann, M. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science 2009, 325, 834–840. [Google Scholar] [CrossRef] [Green Version]
  27. Costantini, C.; Ko, M.H.; Jonas, M.C.; Puglielli, L. A reversible form of lysine acetylation in the ER and Golgi lumen controls the molecular stabilization of BACE1. Biochem. J. 2007, 407, 383–395. [Google Scholar] [CrossRef] [Green Version]
  28. Farrugia, M.A.; Puglielli, L. Nepsilon-lysine acetylation in the endoplasmic reticulum—A novel cellular mechanism that regulates proteostasis and autophagy. J. Cell Sci. 2018, 131. [Google Scholar] [CrossRef] [Green Version]
  29. Wagner, G.R.; Hirschey, M.D. Nonenzymatic Protein Acylation as a Carbon Stress Regulated by Sirtuin Deacylases. Mol. Cell 2014, 54, 5–16. [Google Scholar] [CrossRef] [Green Version]
  30. Ghanta, S.; Grossmann, R.E.; 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, 48, 561–574. [Google Scholar] [CrossRef] [Green Version]
  31. Allfrey, V.G.; Faulkner, R.; Mirsky, A.E. Acetylation and Methylation of Histones and Their Possible Role in the Regulation of Rna Synthesis. Proc. Natl. Acad. Sci. USA 1964, 51, 786–794. [Google Scholar] [CrossRef] [Green Version]
  32. Choudhary, C.; Weinert, B.T.; Nishida, Y.; Verdin, E.; Mann, M. The growing landscape of lysine acetylation links metabolism and cell signalling. Nat. Rev. Mol. Cell Biol. 2014, 15, 536–550. [Google Scholar] [CrossRef] [PubMed]
  33. Soppa, J. Protein acetylation in archaea, bacteria, and eukaryotes. Archaea 2010, 2010. [Google Scholar] [CrossRef] [PubMed]
  34. Paik, W.K.; Pearson, D.; Lee, H.W.; Kim, S. Nonenzymatic acetylation of histones with acetyl-CoA. Biochim. Biophys. Acta 1970, 213, 513–522. [Google Scholar] [CrossRef]
  35. Wagner, G.R.; Payne, R.M. Widespread and enzyme-independent Nepsilon-acetylation and Nepsilon-succinylation of proteins in the chemical conditions of the mitochondrial matrix. J. Biol. Chem. 2013, 288, 29036–29045. [Google Scholar] [CrossRef] [Green Version]
  36. Allfrey, V.G.; Mirsky, A.E. Structural Modifications of Histones and their Possible Role in the Regulation of RNA Synthesis. Science 1964, 144, 559. [Google Scholar] [CrossRef] [PubMed]
  37. Baddiley, J.; Kekwick, R.A.; Thain, E.M. A new method for acetylating proteins. Nature 1952, 170, 968–970. [Google Scholar] [CrossRef]
  38. Weinert, B.T.; Iesmantavicius, V.; Moustafa, T.; Scholz, C.; Wagner, S.A.; Magnes, C.; Zechner, R.; Choudhary, C. Acetylation dynamics and stoichiometry in Saccharomyces cerevisiae. Mol. Syst. Biol. 2015, 11, 833. [Google Scholar] [CrossRef]
  39. Weinert, B.T.; Iesmantavicius, V.; Moustafa, T.; Scholz, C.; Wagner, S.A.; Magnes, C.; Zechner, R.; Choudhary, C. Acetylation dynamics and stoichiometry in Saccharomyces cerevisiae. Mol. Syst. Biol. 2014, 10, 716. [Google Scholar] [CrossRef]
  40. Casey, J.R.; Grinstein, S.; Orlowski, J. Sensors and regulators of intracellular pH. Nat. Rev. Mol. Cell Biol. 2010, 11, 50–61. [Google Scholar] [CrossRef]
  41. Garland, P.B.; Shepherd, D.; Yates, D.W. Steady-state concentrations of coenzyme A, acetyl-coenzyme A and long-chain fatty acyl-coenzyme A in rat-liver mitochondria oxidizing palmitate. Biochem. J. 1965, 97, 587–594. [Google Scholar] [CrossRef] [Green Version]
  42. D’Alayer, J.; Expert-Bezancon, N.; Beguin, P. Time- and temperature-dependent acetylation of the chemokine RANTES produced in recombinant Escherichia coli. Protein Expr. Purif. 2007, 55, 9–16. [Google Scholar] [CrossRef] [PubMed]
  43. Hentschel, A.; Zahedi, R.P.; Ahrends, R. Protein lipid modifications-More than just a greasy ballast. Proteomics 2016, 16, 759–782. [Google Scholar] [CrossRef] [PubMed]
  44. Lanyon-Hogg, T.; Faronato, M.; Serwa, R.A.; Tate, E.W. Dynamic Protein Acylation: New Substrates, Mechanisms, and Drug Targets. Trends Biochem. Sci. 2017, 42, 566–581. [Google Scholar] [CrossRef] [PubMed]
  45. Chamberlain, L.H.; Shipston, M.J. The physiology of protein S-acylation. Physiol. Rev. 2015, 95, 341–376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Fukata, M.; Fukata, Y.; Adesnik, H.; Nicoll, R.A.; Bredt, D.S. Identification of PSD-95 palmitoylating enzymes. Neuron 2004, 44, 987–996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Roth, A.F.; Feng, Y.; Chen, L.; Davis, N.G. The yeast DHHC cysteine-rich domain protein Akr1p is a palmitoyl transferase. J. Cell Biol. 2002, 159, 23–28. [Google Scholar] [CrossRef] [Green Version]
  48. Politis, E.G.; Roth, A.F.; Davis, N.G. Transmembrane topology of the protein palmitoyl transferase Akr1. J. Biol. Chem. 2005, 280, 10156–10163. [Google Scholar] [CrossRef] [Green Version]
  49. Narita, T.; Weinert, B.T.; Choudhary, C. Functions and mechanisms of non-histone protein acetylation. Nat. Rev. Mol. Cell Biol. 2019, 20, 156–174. [Google Scholar] [CrossRef]
  50. Cohen, T.J.; Hwang, A.W.; Restrepo, C.R.; Yuan, C.X.; Trojanowski, J.Q.; Lee, V.M. An acetylation switch controls TDP-43 function and aggregation propensity. Nat. Commun. 2015, 6, 5845. [Google Scholar] [CrossRef] [Green Version]
  51. Min, S.W.; Chen, X.; Tracy, T.E.; Li, Y.; Zhou, Y.; Wang, C.; Shirakawa, K.; Minami, S.S.; Defensor, E.; Mok, S.A.; et al. Critical role of acetylation in tau-mediated neurodegeneration and cognitive deficits. Nat. Med. 2015, 21, 1154–1162. [Google Scholar] [CrossRef] [Green Version]
  52. Clayton-Smith, J.; O’ullivan, J.; Daly, S.; Bhaskar, S.; Day, R.; Anderson, B.; Voss, A.K.; Thomas, T.; Biesecker, L.G.; Smith, P.; et al. Whole-exome-sequencing identifies mutations in histone acetyltransferase gene KAT6B in individuals with the Say-Barber-Biesecker variant of Ohdo syndrome. Am. J. Hum. Genet. 2011, 89, 675–681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Simpson, M.A.; Deshpande, C.; Dafou, D.; Vissers, L.E.; Woollard, W.J.; Holder, S.E.; Gillessen-Kaesbach, G.; Derks, R.; White, S.M.; Cohen-Snuijf, R.; et al. De novo mutations of the gene encoding the histone acetyltransferase KAT6B cause Genitopatellar syndrome. Am. J. Hum. Genet. 2012, 90, 290–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Campeau, P.M.; Kim, J.C.; Lu, J.T.; Schwartzentruber, J.A.; Abdul-Rahman, O.A.; Schlaubitz, S.; Murdock, D.M.; Jiang, M.M.; Lammer, E.J.; Enns, G.M.; et al. Mutations in KAT6B, encoding a histone acetyltransferase, cause Genitopatellar syndrome. Am. J. Hum. Genet. 2012, 90, 282–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Tham, E.; Lindstrand, A.; Santani, A.; Malmgren, H.; Nesbitt, A.; Dubbs, H.A.; Zackai, E.H.; Parker, M.J.; Millan, F.; Rosenbaum, K.; et al. Dominant mutations in KAT6A cause intellectual disability with recognizable syndromic features. Am. J. Hum. Genet. 2015, 96, 507–513. [Google Scholar] [CrossRef] [Green Version]
  56. Arboleda, V.A.; Lee, H.; Dorrani, N.; Zadeh, N.; Willis, M.; Macmurdo, C.F.; Manning, M.A.; Kwan, A.; Hudgins, L.; Barthelemy, F.; et al. De novo nonsense mutations in KAT6A, a lysine acetyl-transferase gene, cause a syndrome including microcephaly and global developmental delay. Am. J. Hum. Genet. 2015, 96, 498–506. [Google Scholar] [CrossRef] [Green Version]
  57. Pasqualucci, L.; Dominguez-Sola, D.; Chiarenza, A.; Fabbri, G.; Grunn, A.; Trifonov, V.; Kasper, L.H.; Lerach, S.; Tang, H.; Ma, J.; et al. Inactivating mutations of acetyltransferase genes in B-cell lymphoma. Nature 2011, 471, 189–195. [Google Scholar] [CrossRef] [Green Version]
  58. Houtkooper, R.H.; Pirinen, E.; Auwerx, J. Sirtuins as regulators of metabolism and healthspan. Nat. Rev. Mol. Cell Biol. 2012, 13, 225–238. [Google Scholar] [CrossRef] [Green Version]
  59. Lasko, L.M.; Jakob, C.G.; Edalji, R.P.; Qiu, W.; Montgomery, D.; Digiammarino, E.L.; Hansen, T.M.; Risi, R.M.; Frey, R.; Manaves, V.; et al. Discovery of a selective catalytic p300/CBP inhibitor that targets lineage-specific tumours. Nature 2017, 550, 128–132. [Google Scholar] [CrossRef]
  60. Baell, J.B.; Leaver, D.J.; Hermans, S.J.; Kelly, G.L.; Brennan, M.S.; Downer, N.L.; Nguyen, N.; Wichmann, J.; McRae, H.M.; Yang, Y.; et al. Inhibitors of histone acetyltransferases KAT6A/B induce senescence and arrest tumour growth. Nature 2018, 560, 253–257. [Google Scholar] [CrossRef]
  61. Salaun, C.; Greaves, J.; Chamberlain, L.H. The intracellular dynamic of protein palmitoylation. J. Cell Biol. 2010, 191, 1229–1238. [Google Scholar] [CrossRef] [Green Version]
  62. Liu, P.; Jiao, B.; Zhang, R.; Zhao, H.; Zhang, C.; Wu, M.; Li, D.; Zhao, X.; Qiu, Q.; Li, J.; et al. Palmitoylacyltransferase Zdhhc9 inactivation mitigates leukemogenic potential of oncogenic Nras. Leukemia 2016, 30, 1225–1228. [Google Scholar] [CrossRef] [PubMed]
  63. Jiang, Y.-X.; Du, Z.-M.; Jiao, L.; Shao, Q.; Fu, S.; Shao, J.-Y.; Zhu, X.-F.; Ernberg, I.; Li, Y.-H. Inhibition of MiR-155 suppresses cell migration in nasopharyngeal carcinoma through targeting ZDHHC2. Int. J. Clin. Exp. Med. 2015, 8, 8472–8484. [Google Scholar] [PubMed]
  64. Cho, E.; Park, M. Palmitoylation in Alzheimer’s disease and other neurodegenerative diseases. Pharmacol. Res. 2016, 111, 133–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Carafa, V.; Nebbioso, A.; Altucci, L. Sirtuins and disease: The road ahead. Front. Pharm. 2012, 3, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zhao, S.; Xu, W.; Jiang, W.; Yu, W.; Lin, Y.; Zhang, T.; Yao, J.; Zhou, L.; Zeng, Y.; Li, H.; et al. Regulation of cellular metabolism by protein lysine acetylation. Science 2010, 327, 1000–1004. [Google Scholar] [CrossRef] [Green Version]
  67. Finkel, T.; Deng, C.X.; Mostoslavsky, R. Recent progress in the biology and physiology of sirtuins. Nature 2009, 460, 587–591. [Google Scholar] [CrossRef] [Green Version]
  68. Brownlee, M. Biochemistry and molecular cell biology of diabetic complications. Nature 2001, 414, 813–820. [Google Scholar] [CrossRef]
  69. Fukami, K.; Yamagishi, S.; Okuda, S. Role of AGEs-RAGE system in cardiovascular disease. Curr. Pharm. Des. 2014, 20, 2395–2402. [Google Scholar] [CrossRef]
  70. Houtkooper, R.H.; Mouchiroud, L.; Ryu, D.; Moullan, N.; Katsyuba, E.; Knott, G.; Williams, R.W.; Auwerx, J. Mitonuclear protein imbalance as a conserved longevity mechanism. Nature 2013, 497, 451–457. [Google Scholar] [CrossRef] [Green Version]
  71. Gottlieb, S.; Esposito, R.E. A new role for a yeast transcriptional silencer gene, SIR2, in regulation of recombination in ribosomal DNA. Cell 1989, 56, 771–776. [Google Scholar] [CrossRef]
  72. Bennett, C.B.; Snipe, J.R.; Westmoreland, J.W.; Resnick, M.A. SIR functions are required for the toleration of an unrepaired double-strand break in a dispensable yeast chromosome. Mol. Cell. Biol. 2001, 21, 5359–5373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Brachmann, C.B.; Sherman, J.M.; Devine, S.E.; Cameron, E.E.; Pillus, L.; Boeke, J.D. The SIR2 gene family, conserved from bacteria to humans, functions in silencing, cell cycle progression, and chromosome stability. Genes Dev. 1995, 9, 2888–2902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Araki, T.; Sasaki, Y.; Milbrandt, J. Increased nuclear NAD biosynthesis and SIRT1 activation prevent axonal degeneration. Science 2004, 305, 1010–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Picard, F.; Kurtev, M.; Chung, N.; Topark-Ngarm, A.; Senawong, T.; Machado De Oliveira, R.; Leid, M.; McBurney, M.W.; Guarente, L. Sirt1 promotes fat mobilization in white adipocytes by repressing PPAR-gamma. Nature 2004, 429, 771–776. [Google Scholar] [CrossRef]
  76. Brunet, A.; Sweeney, L.B.; Sturgill, J.F.; Chua, K.F.; Greer, P.L.; Lin, Y.; Tran, H.; Ross, S.E.; Mostoslavsky, R.; Cohen, H.Y.; et al. Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science 2004, 303, 2011–2015. [Google Scholar] [CrossRef] [Green Version]
  77. Luo, J.; Nikolaev, A.Y.; Imai, S.; Chen, D.; Su, F.; Shiloh, A.; Guarente, L.; Gu, W. Negative control of p53 by Sir2alpha promotes cell survival under stress. Cell 2001, 107, 137–148. [Google Scholar] [CrossRef] [Green Version]
  78. Lin, S.J.; Defossez, P.A.; Guarente, L. Requirement of NAD and SIR2 for life-span extension by calorie restriction in Saccharomyces cerevisiae. Science 2000, 289, 2126–2128. [Google Scholar] [CrossRef] [Green Version]
  79. 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. [Google Scholar] [CrossRef] [Green Version]
  80. Du, J.; Zhou, Y.; Su, X.; Yu, J.J.; Khan, S.; Jiang, H.; Kim, J.; Woo, J.; Kim, J.H.; Choi, B.H.; et al. Sirt5 is a NAD-dependent protein lysine demalonylase and desuccinylase. Science 2011, 334, 806–809. [Google Scholar] [CrossRef] [Green Version]
  81. 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. Mol. Cell. Proteom. 2011, 10. [Google Scholar] [CrossRef] [Green Version]
  82. Tan, M.; Peng, C.; Anderson, K.A.; Chhoy, P.; Xie, Z.; Dai, L.; Park, J.; Chen, Y.; Huang, H.; Zhang, Y.; et al. Lysine glutarylation is a protein posttranslational modification regulated by SIRT5. Cell Metab. 2014, 19, 605–617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Anderson, K.A.; Huynh, F.K.; Fisher-Wellman, K.; Stuart, J.D.; Peterson, B.S.; Douros, J.D.; Wagner, G.R.; Thompson, J.W.; Madsen, A.S.; Green, M.F.; et al. SIRT4 Is a Lysine Deacylase that Controls Leucine Metabolism and Insulin Secretion. Cell Metab. 2017, 25, 838–855.e15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Haberland, M.; Montgomery, R.L.; Olson, E.N. The many roles of histone deacetylases in development and physiology: Implications for disease and therapy. Nat. Rev. Genet. 2009, 10, 32–42. [Google Scholar] [CrossRef] [PubMed]
  85. Jones, P.; Altamura, S.; De Francesco, R.; Gallinari, P.; Lahm, A.; Neddermann, P.; Rowley, M.; Serafini, S.; Steinkuhler, C. Probing the elusive catalytic activity of vertebrate class IIa histone deacetylases. Bioorg Med. Chem. Lett. 2008, 18, 1814–1819. [Google Scholar] [CrossRef] [PubMed]
  86. Yang, X.J.; Seto, E. The Rpd3/Hda1 family of lysine deacetylases: From bacteria and yeast to mice and men. Nat. Rev. Mol. Cell Biol. 2008, 9, 206–218. [Google Scholar] [CrossRef] [Green Version]
  87. Jackson, M.D.; Denu, J.M. Structural identification of 2′- and 3′-O-acetyl-ADP-ribose as novel metabolites derived from the Sir2 family of beta -NAD+-dependent histone/protein deacetylases. J. Biol. Chem. 2002, 277, 18535–18544. [Google Scholar] [CrossRef] [Green Version]
  88. Rine, J.; Herskowitz, I. Four genes responsible for a position effect on expression from HML and HMR in Saccharomyces cerevisiae. Genetics 1987, 116, 9–22. [Google Scholar]
  89. Frye, R.A. Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem. Biophys. Res. Commun. 2000, 273, 793–798. [Google Scholar] [CrossRef]
  90. Sanders, B.D.; Jackson, B.; Marmorstein, R. Structural basis for sirtuin function: What we know and what we don’t. Biochim. Biophys. Acta 2010, 1804, 1604–1616. [Google Scholar] [CrossRef] [Green Version]
  91. Hawse, W.F.; Wolberger, C. Structure-based mechanism of ADP-ribosylation by sirtuins. J. Biol. Chem. 2009, 284, 33654–33661. [Google Scholar] [CrossRef] [Green Version]
  92. Tanner, K.G.; Landry, J.; Sternglanz, R.; Denu, J.M. Silent information regulator 2 family of NAD- dependent histone/protein deacetylases generates a unique product, 1-O-acetyl-ADP-ribose. Proc. Natl. Acad. Sci. USA 2000, 97, 14178–14182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Bheda, P.; Jing, H.; Wolberger, C.; Lin, H. The Substrate Specificity of Sirtuins. Ann. Rev. Biochem. 2016, 85, 405–429. [Google Scholar] [CrossRef] [PubMed]
  94. Avalos, J.L.; Celic, I.; Muhammad, S.; Cosgrove, M.S.; Boeke, J.D.; Wolberger, C. Structure of a Sir2 enzyme bound to an acetylated p53 peptide. Mol. Cell 2002, 10, 523–535. [Google Scholar] [CrossRef]
  95. Avalos, J.L.; Boeke, J.D.; Wolberger, C. Structural basis for the mechanism and regulation of Sir2 enzymes. Mol. Cell 2004, 13, 639–648. [Google Scholar] [CrossRef]
  96. Min, J.; Landry, J.; Sternglanz, R.; Xu, R.M. Crystal structure of a SIR2 homolog-NAD complex. Cell 2001, 105, 269–279. [Google Scholar] [CrossRef] [Green Version]
  97. Hoff, K.G.; Avalos, J.L.; Sens, K.; Wolberger, C. Insights into the Sirtuin Mechanism from Ternary Complexes Containing NAD+ and Acetylated Peptide. Structure 2006, 14, 1231–1240. [Google Scholar] [CrossRef] [Green Version]
  98. Rauh, D.; Fischer, F.; Gertz, M.; Lakshminarasimhan, M.; Bergbrede, T.; Aladini, F.; Kambach, C.; Becker, C.F.; Zerweck, J.; Schutkowski, M.; et al. An acetylome peptide microarray reveals specificities and deacetylation substrates for all human sirtuin isoforms. Nat. Commun. 2013, 4, 2327. [Google Scholar] [CrossRef] [Green Version]
  99. De Boor, S.; Knyphausen, P.; Kuhlmann, N.; Wroblowski, S.; Brenig, J.; Scislowski, L.; Baldus, L.; Nolte, H.; Kruger, M.; Lammers, M. Small GTP-binding protein Ran is regulated by posttranslational lysine acetylation. Proc. Natl. Acad. Sci. USA 2015, 112, E3679–E3688. [Google Scholar] [CrossRef] [Green Version]
  100. Kuhlmann, N.; Wroblowski, S.; Knyphausen, P.; de Boor, S.; Brenig, J.; Zienert, A.Y.; Meyer-Teschendorf, K.; Praefcke, G.J.; Nolte, H.; Kruger, M.; et al. Structural and Mechanistic Insights into the Regulation of the Fundamental Rho Regulator RhoGDIalpha by Lysine Acetylation. J. Biol. Chem. 2016, 291, 5484–5499. [Google Scholar] [CrossRef] [Green Version]
  101. Zhao, K.; Chai, X.; Marmorstein, R. Structure and substrate binding properties of cobB, a Sir2 homolog protein deacetylase from Escherichia coli. J. Mol. Biol. 2004, 337, 731–741. [Google Scholar] [CrossRef]
  102. Tanno, M.; Sakamoto, J.; Miura, T.; Shimamoto, K.; Horio, Y. Nucleocytoplasmic Shuttling of the NAD+-dependent Histone Deacetylase SIRT1. J. Biol. Chem. 2007, 282, 6823–6832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Gomes, P.; Fleming Outeiro, T.; Cavadas, C. Emerging Role of Sirtuin 2 in the Regulation of Mammalian Metabolism. Trends Pharmacol. Sci. 2015, 36, 756–768. [Google Scholar] [CrossRef] [PubMed]
  104. Ansari, A.; Rahman, M.S.; Saha, S.K.; Saikot, F.K.; Deep, A.; Kim, K.H. Function of the SIRT3 mitochondrial deacetylase in cellular physiology, cancer, and neurodegenerative disease. Aging Cell 2017, 16, 4–16. [Google Scholar] [CrossRef] [PubMed]
  105. Michan, S.; Sinclair, D. Sirtuins in mammals: Insights into their biological function. Biochem. J. 2007, 404, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Nishida, Y.; Rardin, M.J.; Carrico, C.; He, W.; Sahu, A.K.; Gut, P.; Najjar, R.; Fitch, M.; Hellerstein, M.; Gibson, B.W.; et al. SIRT5 Regulates both Cytosolic and Mitochondrial Protein Malonylation with Glycolysis as a Major Target. Mol. Cell 2015, 59, 321–332. [Google Scholar] [CrossRef] [Green Version]
  107. Lin, H. The Enzymatic Activities of Sirtuins. In Introductory Review on Sirtuins in Biology, Aging, and Disease; Academic Press: Cambridge, MA, USA, 2018; pp. 45–62. [Google Scholar] [CrossRef]
  108. Michishita, E.; Park, J.Y.; Burneskis, J.M.; Barrett, J.C.; Horikawa, I. Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Mol. Biol. Cell 2005, 16, 4623–4635. [Google Scholar] [CrossRef] [Green Version]
  109. Feldman, J.L.; Baeza, J.; Denu, J.M. Activation of the protein deacetylase SIRT6 by long-chain fatty acids and widespread deacylation by mammalian sirtuins. J. Biol. Chem. 2013, 288, 31350–31356. [Google Scholar] [CrossRef] [Green Version]
  110. 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. [Google Scholar] [CrossRef] [Green Version]
  111. Li, L.; Shi, L.; Yang, S.; Yan, R.; Zhang, D.; Yang, J.; He, L.; Li, W.; Yi, X.; Sun, L.; et al. SIRT7 is a histone desuccinylase that functionally links to chromatin compaction and genome stability. Nat. Commun. 2016, 7, 12235. [Google Scholar] [CrossRef] [Green Version]
  112. Wei, L.; Hu, F.; Shen, Y.; Chen, Z.; Yu, Y.; Lin, C.C.; Wang, M.C.; Min, W. Live-cell imaging of alkyne-tagged small biomolecules by stimulated Raman scattering. Nat. Methods 2014, 11, 410–412. [Google Scholar] [CrossRef] [Green Version]
  113. Khan, A.N.; Lewis, P.N. Unstructured conformations are a substrate requirement for the Sir2 family of NAD-dependent protein deacetylases. J. Biol. Chem. 2005, 280, 36073–36078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Knyphausen, P.; de Boor, S.; Kuhlmann, N.; Scislowski, L.; Extra, A.; Baldus, L.; Schacherl, M.; Baumann, U.; Neundorf, I.; Lammers, M. Insights into Lysine Deacetylation of Natively Folded Substrate Proteins by Sirtuins. J. Biol. Chem. 2016, 291, 14677–14694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Cosgrove, M.S.; Bever, K.; Avalos, J.L.; Muhammad, S.; Zhang, X.; Wolberger, C. The structural basis of sirtuin substrate affinity. Biochemistry 2006, 45, 7511–7521. [Google Scholar] [CrossRef] [PubMed]
  116. Vaziri, H.; Dessain, S.K.; Eaton, E.N.; Imai, S.-I.; Frye, R.A.; Pandita, T.K.; Guarente, L.; Weinberg, R.A. hSIR2SIRT1 Functions as an NAD-Dependent p53 Deacetylase. Cell 2001, 107, 149–159. [Google Scholar] [CrossRef] [Green Version]
  117. Schwer, B.; Bunkenborg, J.; Verdin, R.O.; Andersen, J.S.; Verdin, E. Reversible lysine acetylation controls the activity of the mitochondrial enzyme acetyl-CoA synthetase 2. Proc. Natl. Acad. Sci. USA 2006, 103, 10224–10229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Kawahara, T.L.A.; Michishita, E.; Adler, A.S.; Damian, M.; Berber, E.; Lin, M.; McCord, R.A.; Ongaigui, K.C.L.; Boxer, L.D.; Chang, H.Y.; et al. SIRT6 Links Histone H3 Lysine 9 Deacetylation to NF-κB-Dependent Gene Expression and Organismal Life Span. Cell 2009, 136, 62–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Zhong, L.; D’Urso, A.; Toiber, D.; Sebastian, C.; Henry, R.E.; Vadysirisack, D.D.; Guimaraes, A.; Marinelli, B.; Wikstrom, J.D.; Nir, T.; et al. The Histone Deacetylase Sirt6 Regulates Glucose Homeostasis via Hif1α. Cell 2010, 140, 280–293. [Google Scholar] [CrossRef] [Green Version]
  120. Barber, M.F.; Michishita-Kioi, E.; Xi, Y.; Tasselli, L.; Kioi, M.; Moqtaderi, Z.; Tennen, R.I.; Paredes, S.; Young, N.L.; Chen, K.; et al. SIRT7 links H3K18 deacetylation to maintenance of oncogenic transformation. Nature 2012, 487, 114–118. [Google Scholar] [CrossRef]
  121. Sauve, A.A. Sirtuin chemical mechanisms. Biochim. Biophys. Acta 2010, 1804, 1591–1603. [Google Scholar] [CrossRef] [Green Version]
  122. Shi, Y.; Zhou, Y.; Wang, S.; Zhang, Y. Sirtuin Deacetylation Mechanism and Catalytic Role of the Dynamic Cofactor Binding Loop. J. Phys. Chem. Lett. 2013, 4, 491–495. [Google Scholar] [CrossRef] [Green Version]
  123. Zhou, Y.; Zhang, H.; He, B.; Du, J.; Lin, H.; Cerione, R.A.; Hao, Q. The bicyclic intermediate structure provides insights into the desuccinylation mechanism of human sirtuin 5 (SIRT5). J. Biol. Chem. 2012, 287, 28307–28314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Sauve, A.A.; Celic, I.; Avalos, J.; Deng, H.; Boeke, J.D.; Schramm, V.L. Chemistry of gene silencing: The mechanism of NAD+-dependent deacetylation reactions. Biochemistry 2001, 40, 15456–15463. [Google Scholar] [CrossRef] [PubMed]
  125. Tanny, J.C.; Moazed, D. Coupling of histone deacetylation to NAD breakdown by the yeast silencing protein Sir2: Evidence for acetyl transfer from substrate to an NAD breakdown product. Proc. Natl. Acad. Sci. USA 2001, 98, 415–420. [Google Scholar] [CrossRef] [PubMed]
  126. Bosch-Presegue, L.; Vaquero, A. Sirtuin-dependent epigenetic regulation in the maintenance of genome integrity. FEBS J. 2015, 282, 1745–1767. [Google Scholar] [CrossRef]
  127. Masri, S.; Sassone-Corsi, P. Sirtuins and the circadian clock: Bridging chromatin and metabolism. Sci. Signal. 2014, 7, re6. [Google Scholar] [CrossRef]
  128. Guarente, L. Sirtuins, aging, and metabolism. Cold Spring Harb Symp. Quant. Biol. 2011, 76, 81–90. [Google Scholar] [CrossRef]
  129. Haigis, M.C.; Sinclair, D.A. Mammalian sirtuins: Biological insights and disease relevance. Annu. Rev. Pathol. 2010, 5, 253–295. [Google Scholar] [CrossRef] [Green Version]
  130. Longo, V.D.; Antebi, A.; Bartke, A.; Barzilai, N.; Brown-Borg, H.M.; Caruso, C.; Curiel, T.J.; Cabo, R.; Franceschi, C.; Gems, D.; et al. Interventions to Slow Aging in Humans: Are We Ready? Aging Cell 2015, 14, 497–510. [Google Scholar] [CrossRef]
  131. Watroba, M.; Szukiewicz, D. The role of sirtuins in aging and age-related diseases. Adv. Med. Sci. 2016, 61, 52–62. [Google Scholar] [CrossRef]
  132. Chen, D.; Steele, A.D.; Lindquist, S.; Guarente, L. Increase in activity during calorie restriction requires Sirt1. Science 2005, 310, 1641. [Google Scholar] [CrossRef] [Green Version]
  133. Boily, G.; Seifert, E.L.; Bevilacqua, L.; He, X.H.; Sabourin, G.; Estey, C.; Moffat, C.; Crawford, S.; Saliba, S.; Jardine, K.; et al. SirT1 regulates energy metabolism and response to caloric restriction in mice. PLoS ONE 2008, 3, e1759. [Google Scholar] [CrossRef]
  134. Grabowska, W.; Sikora, E.; Bielak-Zmijewska, A. Sirtuins, a promising target in slowing down the ageing process. Biogerontology 2017, 18, 447–476. [Google Scholar] [CrossRef] [Green Version]
  135. Chung, J.H.; Manganiello, V.; Dyck, J.R.B. Resveratrol as a calorie restriction mimetic: Therapeutic implications. Trends Cell Biol. 2012, 22, 546–554. [Google Scholar] [CrossRef] [Green Version]
  136. Bosch-Presegue, L.; Vaquero, A. The dual role of sirtuins in cancer. Genes Cancer 2011, 2, 648–662. [Google Scholar] [CrossRef]
  137. Deng, C.-X. SIRT1, Is It a Tumor Promoter or Tumor Suppressor? Int. J. Biol. Sci. 2009, 5, 147–152. [Google Scholar] [CrossRef] [Green Version]
  138. El-Rifai, W.; Zhang, T.; Rong, N.; Chen, J.; Zou, C.; Jing, H.; Zhu, X.; Zhang, W. SIRT1 Expression Is Associated with the Chemotherapy Response and Prognosis of Patients with Advanced NSCLC. PLoS ONE 2013, 8, e79162. [Google Scholar] [CrossRef]
  139. McGlynn, L.M.; Zino, S.; MacDonald, A.I.; Curle, J.; Reilly, J.E.; Mohammed, Z.M.A.; McMillan, D.C.; Mallon, E.; Payne, A.P.; Edwards, J.; et al. SIRT2: Tumour suppressor or tumour promoter in operable breast cancer? Eur. J. Cancer 2014, 50, 290–301. [Google Scholar] [CrossRef]
  140. Ozden, O.; Park, S.-H.; Wagner, B.A.; Yong Song, H.; Zhu, Y.; Vassilopoulos, A.; Jung, B.; Buettner, G.R.; Gius, D. SIRT3 deacetylates and increases pyruvate dehydrogenase activity in cancer cells. Free Radic. Biol. Med. 2014, 76, 163–172. [Google Scholar] [CrossRef] [Green Version]
  141. Miyo, M.; Yamamoto, H.; Konno, M.; Colvin, H.; Nishida, N.; Koseki, J.; Kawamoto, K.; Ogawa, H.; Hamabe, A.; Uemura, M.; et al. Tumour-suppressive function of SIRT4 in human colorectal cancer. Br. J. Cancer 2015, 113, 492–499. [Google Scholar] [CrossRef] [Green Version]
  142. Lu, W.; Zuo, Y.; Feng, Y.; Zhang, M. SIRT5 facilitates cancer cell growth and drug resistance in non-small cell lung cancer. Tumour. Biol. 2014, 35, 10699–10705. [Google Scholar] [CrossRef]
  143. Desantis, V.; Lamanuzzi, A.; Vacca, A. The role of SIRT6 in tumors. Haematologica 2018, 103, 1–4. [Google Scholar] [CrossRef] [Green Version]
  144. Elibol, B.; Kilic, U. High Levels of SIRT1 Expression as a Protective Mechanism Against Disease-Related Conditions. Front. Endocrinol. 2018, 9, 614. [Google Scholar] [CrossRef]
  145. Lutz, M.I.; Milenkovic, I.; Regelsberger, G.; Kovacs, G.G. Distinct patterns of sirtuin expression during progression of Alzheimer’s disease. Neuromol. Med. 2014, 16, 405–414. [Google Scholar] [CrossRef]
  146. Chan, S.H.; Hung, C.H.; Shih, J.Y.; Chu, P.M.; Cheng, Y.H.; Lin, H.C.; Tsai, K.L. SIRT1 inhibition causes oxidative stress and inflammation in patients with coronary artery disease. Redox Biol. 2017, 13, 301–309. [Google Scholar] [CrossRef]
  147. Costa, C.d.S.; Hammes, T.O.; Rohden, F.; Margis, R.; Bortolotto, J.W.; Padoin, A.V.; Mottin, C.C.; Guaragna, R.M. SIRT1 Transcription Is Decreased in Visceral Adipose Tissue of Morbidly Obese Patients with Severe Hepatic Steatosis. Obes. Surg. 2009, 20, 633–639. [Google Scholar] [CrossRef]
  148. Braidy, N.; Jayasena, T.; Poljak, A.; Sachdev, P.S. Sirtuins in cognitive ageing and Alzheimer’s disease. Curr. Opin. Psychiatry 2012, 25, 226–230. [Google Scholar] [CrossRef]
  149. Singh, P.; Hanson, P.S.; Morris, C.M. SIRT1 ameliorates oxidative stress induced neural cell death and is down-regulated in Parkinson’s disease. BMC Neurosci. 2017, 18, 1–13. [Google Scholar] [CrossRef]
  150. Dai, H.; Sinclair, D.A.; Ellis, J.L.; Steegborn, C. Sirtuin activators and inhibitors: Promises, achievements, and challenges. Pharmacol. Ther. 2018, 188, 140–154. [Google Scholar] [CrossRef]
  151. Dompierre, J.P.; Godin, J.D.; Charrin, B.C.; Cordelieres, F.P.; King, S.J.; Humbert, S.; Saudou, F. Histone deacetylase 6 inhibition compensates for the transport deficit in Huntington’s disease by increasing tubulin acetylation. J. Neurosci. 2007, 27, 3571–3583. [Google Scholar] [CrossRef]
  152. Govindarajan, N.; Rao, P.; Burkhardt, S.; Sananbenesi, F.; Schluter, O.M.; Bradke, F.; Lu, J.; Fischer, A. Reducing HDAC6 ameliorates cognitive deficits in a mouse model for Alzheimer’s disease. EMBO Mol. Med. 2013, 5, 52–63. [Google Scholar] [CrossRef]
  153. Saba, N.F.; Magliocca, K.R.; Kim, S.; Muller, S.; Chen, Z.; Owonikoko, T.K.; Sarlis, N.J.; Eggers, C.; Phelan, V.; Grist, W.J.; et al. Acetylated tubulin (AT) as a prognostic marker in squamous cell carcinoma of the head and neck. Head Neck Pathol. 2014, 8, 66–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Boggs, A.E.; Vitolo, M.I.; Whipple, R.A.; Charpentier, M.S.; Goloubeva, O.G.; Ioffe, O.B.; Tuttle, K.C.; Slovic, J.; Lu, Y.; Mills, G.B.; et al. alpha-Tubulin acetylation elevated in metastatic and basal-like breast cancer cells promotes microtentacle formation, adhesion, and invasive migration. Cancer Res. 2015, 75, 203–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Li, P.; Ge, J.; Li, H. Lysine acetyltransferases and lysine deacetylases as targets for cardiovascular disease. Nat. Rev. Cardiol. 2019, 17, 96–115. [Google Scholar] [CrossRef] [PubMed]
  156. McLendon, P.M.; Ferguson, B.S.; Osinska, H.; Bhuiyan, M.S.; James, J.; McKinsey, T.A.; Robbins, J. Tubulin hyperacetylation is adaptive in cardiac proteotoxicity by promoting autophagy. Proc. Natl. Acad. Sci. USA 2014, 111, E5178–E5186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Sinclair, D.A.; Guarente, L. Extrachromosomal rDNA circles--a cause of aging in yeast. Cell 1997, 91, 1033–1042. [Google Scholar] [CrossRef] [Green Version]
  158. Satoh, A.; Stein, L.; Imai, S. The role of mammalian sirtuins in the regulation of metabolism, aging, and longevity. Handb. Exp. Pharm. 2011, 206, 125–162. [Google Scholar] [CrossRef] [Green Version]
  159. Chakraborty, C.; Doss, C.G. Sirtuins family--recent development as a drug target for aging, metabolism, and age related diseases. Curr. Drug Targets 2013, 14, 666–675. [Google Scholar] [CrossRef]
  160. Kane, A.E.; Sinclair, D.A. Sirtuins and NAD(+) in the Development and Treatment of Metabolic and Cardiovascular Diseases. Circ. Res. 2018, 123, 868–885. [Google Scholar] [CrossRef]
  161. Satoh, A.; Brace, C.S.; Rensing, N.; Cliften, P.; Wozniak, D.F.; Herzog, E.D.; Yamada, K.A.; Imai, S. Sirt1 extends life span and delays aging in mice through the regulation of Nk2 homeobox 1 in the DMH and LH. Cell Metab. 2013, 18, 416–430. [Google Scholar] [CrossRef] [Green Version]
  162. Kanfi, Y.; Naiman, S.; Amir, G.; Peshti, V.; Zinman, G.; Nahum, L.; Bar-Joseph, Z.; Cohen, H.Y. The sirtuin SIRT6 regulates lifespan in male mice. Nature 2012, 483, 218–221. [Google Scholar] [CrossRef]
  163. Mitchell, S.J.; Martin-Montalvo, A.; Mercken, E.M.; Palacios, H.H.; Ward, T.M.; Abulwerdi, G.; Minor, R.K.; Vlasuk, G.P.; Ellis, J.L.; Sinclair, D.A.; et al. The SIRT1 activator SRT1720 extends lifespan and improves health of mice fed a standard diet. Cell Rep. 2014, 6, 836–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Mercken, E.M.; Mitchell, S.J.; Martin-Montalvo, A.; Minor, R.K.; Almeida, M.; Gomes, A.P.; Scheibye-Knudsen, M.; Palacios, H.H.; Licata, J.J.; Zhang, Y.; et al. SRT2104 extends survival of male mice on a standard diet and preserves bone and muscle mass. Aging Cell 2014, 13, 787–796. [Google Scholar] [CrossRef] [PubMed]
Figure 1. General overview of the cellular events related with the protein acylation process.
Figure 1. General overview of the cellular events related with the protein acylation process.
Ijms 21 08609 g001
Figure 2. Classification of different lysine deacylases (KDACs) based on their phylogenetic conservation and/or sequence similarities.
Figure 2. Classification of different lysine deacylases (KDACs) based on their phylogenetic conservation and/or sequence similarities.
Ijms 21 08609 g002
Figure 3. (a) VMD representation of the crystallographic structure of a ternary complex from a wild-type Sir2Tm enzyme from Thermatoga maritima (PDB ID: 2H4F [97]) (new cartoon representation) with 9 residues from the acetylated p53 peptide (licorice representation), one NAD+ molecule (CPK representation) and one Zn2+ molecule (VDW representation). The peptide-binding domain (orange) was obtained from the PDB ID structure 2H59 [97] through structural alignment. (b) VMD representation of the acylated peptide-binding cleft and (c) VMD representation of the C pocket with NAD+ inside it.
Figure 3. (a) VMD representation of the crystallographic structure of a ternary complex from a wild-type Sir2Tm enzyme from Thermatoga maritima (PDB ID: 2H4F [97]) (new cartoon representation) with 9 residues from the acetylated p53 peptide (licorice representation), one NAD+ molecule (CPK representation) and one Zn2+ molecule (VDW representation). The peptide-binding domain (orange) was obtained from the PDB ID structure 2H59 [97] through structural alignment. (b) VMD representation of the acylated peptide-binding cleft and (c) VMD representation of the C pocket with NAD+ inside it.
Ijms 21 08609 g003
Figure 4. Sirtuins’ catalytic mechanism proposal [122]. The red arrows represent the movement of electrons.
Figure 4. Sirtuins’ catalytic mechanism proposal [122]. The red arrows represent the movement of electrons.
Ijms 21 08609 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Teixeira, C.S.S.; Cerqueira, N.M.F.S.A.; Gomes, P.; Sousa, S.F. A Molecular Perspective on Sirtuin Activity. Int. J. Mol. Sci. 2020, 21, 8609. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21228609

AMA Style

Teixeira CSS, Cerqueira NMFSA, Gomes P, Sousa SF. A Molecular Perspective on Sirtuin Activity. International Journal of Molecular Sciences. 2020; 21(22):8609. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21228609

Chicago/Turabian Style

Teixeira, Carla S. S., Nuno M. F. S. A. Cerqueira, Pedro Gomes, and Sérgio F. Sousa. 2020. "A Molecular Perspective on Sirtuin Activity" International Journal of Molecular Sciences 21, no. 22: 8609. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms21228609

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