Open access peer-reviewed chapter

Silent Information Regulator 2 from Trypanosoma cruzi Is a Potential Target to Infection Control

Written By

Luís Gaspar, Terry K. Smith, Nilmar Silvio Moretti, Sergio Schenkman and Anabela Cordeiro-da-Silva

Submitted: 16 May 2017 Reviewed: 06 April 2018 Published: 12 September 2018

DOI: 10.5772/intechopen.77030

From the Edited Volume

Chagas Disease - Basic Investigations and Challenges

Edited by Veeranoot Nissapatorn and Helieh S. Oz

Chapter metrics overview

1,173 Chapter Downloads

View Full Metrics

Abstract

Human trypanosomiasis is a neglected tropical disease caused by protozoan parasites of the genus Trypanosoma. Trypanosoma brucei is responsible for sleeping sickness, also called African trypanosomiasis, while Trypanosoma cruzi causes Chagas disease, or American trypanosomiasis. Together, these diseases are responsible for significant mortality, morbidity and lost productivity in the endemic regions. There are no vaccines and treatments rely on drugs with limited efficacy, high cost, serious side effects and long administration periods. Since these diseases affect mostly the poor, there is no economic interest in the development of new drugs by pharmaceutical companies, and hopes for new treatments rely on public initiatives, public-private partnerships or philanthropic programs. The first step in the discovery of new drugs involves the identification of active molecules as starting points for further development, by either employing whole cells or by specific molecular target screenings. Research efforts undertaken by the authors’ groups have focused on exploiting both strategies in the search for new molecules for trypanosomiasis drug discovery. In this chapter, we focus on Chagas disease and the recently uncovered potential of using sirtuins as targets for infection control.

Keywords

  • Trypanosoma cruzi
  • Chagas disease
  • sirtuins
  • drug discovery
  • chemotherapy

1. Introduction

Despite the efforts of many individuals and organizations over the years, human trypanosomiasis remains one of the most neglected diseases in the world. Chagas disease in particular, is a leading cause of disease and disability in Latin America, with thousands of deaths every year [1]. The negligence is particularly patented by the lack of new drugs. Indeed, the available treatment options, benznidazole and nifurtimox (Figure 1), were discovered more than 40 years ago. Different strategies have been employed to control the disease, but the most impactful so far has been the control of the transmitting vector led by the World Health Organization (WHO) and Pan American Health Organization (PAHO). Vector control has caused the reduction of cases from a staggering 24 million in the 1980s to about 6 million nowadays [1, 2], with some countries considered to be free of domestic vectorial transmission. Continued and rigorous implementation of the disinfestation programs in the remaining zones should decrease even further the global numbers of Trypanosoma cruzi vectorial transmission. Also, the screening of donor blood and transplant organs in endemic regions and other parts of the world has greatly reduced the number of cases transmitted by this route. Whereas the global interruption of the domestic cycle will be a major breakthrough and reduce to a minimum the number of new cases of Chagas disease, complete eradication of the parasite, however, is unlikely to be achieved due to the huge natural reservoir of T. cruzi and the many species of triatomines capable of its transmission in the sylvatic environment [3, 4, 5].

Figure 1.

Various trypanocidal agents.

What then, stands in the way of disease control for Chagas? First, no major technological advances are required to interrupt vectorial transmission responsible for the majority of new cases; second, decades of research in the molecular understanding of T. cruzi biology, the particularities of pathogenesis of the disease or the dynamics of immune response against the parasite have failed to translate into therapeutic alternatives; and finally, vaccination, either preventive or therapeutic, has remained an elusive achievement. The answer lies in new therapeutics. New, safer, cheap, easy to administer and efficacious drugs that are able to treat not only new cases, but also the millions of people already affected by the chronic stage and indeterminate form of the disease. There is growing evidence that chronic manifestations are ultimately related with inflammation resulting from parasite persistence [6, 7] and effective treatment of these cases would be highly beneficial to stop the development of symptomatology.

Active drug discovery efforts for Chagas disease have been restricted, until some years ago, to just a very limited number of groups, mostly based in academia. As a consequence, results have been sporadic, slow, ineffective and highly dependent upon intermittent funding, failing to deliver an alternative treatment. Chagas disease is as much neglected by the pharmaceutical industry as it is by research funding organizations, whose majority of funds are directed to developed world diseases.

Only recently has drug discovery for Chagas been met with concerted, focused efforts. While still not privately embraced by pharmaceutical companies, public-private partnerships have been set up with the objective of bringing together the biology expertise from academia and the technical expertise, facilities and resources from pharmaceutical industries. One organization that is leading the efforts to find new therapies for neglected diseases, including human trypanosomiasis, is the Drugs for Neglected Diseases Initiative (DNDi), that has been involved in coordinating activities from early drug discovery to the launch and conclusion of clinical trials for some candidates like inhibitors of ergosterol biosynthesis of T. cruzi. The Bill and Melinda Gates Foundation, a non-governmental organization devoted to human development in underdeveloped countries, has committed to help control neglected diseases by signing The London Declaration on Neglected Diseases together with the WHO, the World Bank and 13 leading pharmaceutical industries. The Declaration states that by 2020 the signers will achieve, among other ambitious milestones, the eradication of Human African trypanosomiasis (HAT) and the control of American trypanosomiasis. Since then, other initiatives have been launched with the objective of boosting the research in new drugs for Trypanosomal diseases, like the European Commission Seventh Framework Program (FP7) consortia NMTrypI and KINDReD. The recent award of the Nobel Prize in Physiology or Medicine 2015 to William C. Campbell, Satoshi Omura and Youyou Tu is a recognition of the importance of drug discovery for parasitology and should further increase the awareness of neglected diseases by the international community.

Pharmaceutical research and drug discovery for infectious disease have historically began with what would be classified today as phenotypic assays, and can be traced to the pioneering work of Paul Ehrlich in the nineteenth century, while testing the effect of different dyes in trypanosomes [8]. Cultures of the microorganism of interest, bacteria or parasites, were incubated with a compound of interest, and the selective staining of the dyes was monitored by microscopy. Products of such “dye therapy” approaches were in the origin of well-known chemicals like the crystal violet dye that was proposed to be used in blood banks of endemic areas to kill T. cruzi parasites present in transfusion blood as a way to reduce transmission by this route [9]. Another example is trypan blue, that is still widely used as a cell biology reagent and that was the starting point for the design of the colorless analogue Suramin, a drug still in use today for the treatment of HAT and infected animals as well [10]. Such early whole cells drug screening principles were also central to the development of nifurtimox and benznidazole (Figure 1) in the 1960s and 1970s, by the pharmaceutical companies Bayer and Roche, respectively [11].

With the genomic era there was a dramatic shift in the way new drugs are discovered. The past 20 years have witnessed incredible advancements in genomics, proteomics, structural biology, computational chemistry and structure based drug-design, that coupled with high-throughput screening and combinatorial chemistry have helped to shape the reductionist mentality “one gene—one protein—one drug” [12]. However, the complexity of many diseases and the capacity of adaptation to adverse conditions, like the presence of a xenobiotic, evidenced by many living cells, have brought the more naïve phenotypic whole-cell screening strategies back to the spotlight. With modern phenotypic approaches, the effect of a pure molecule in a fully intact whole living organism, bacteria, parasites or human-derived cell lines, results in the identification of hit compounds that are potentially useful as scaffolds for further medicinal chemistry optimization.

While early phenotypic screenings for T. cruzi sometimes used the insect-specific epimastigote stage due of its extracellular nature and ease to culture, the use of reporter genes expressed during the clinically relevant stage of the disease, the intracellular amastigotes, has met a widespread application. The first of such assays was based upon the β-galactosidase-expressing parasites that made possible the detection of anti-T. cruzi activity by a colorimetric reaction [13]. Later, tdt-tomato and luciferase genes were also constitutively expressed in parasites, allowing more sensitive measurement of a fluorescent or luminescent signals, respectively [14].

However, the use of genetically unmodified parasites has always been an attractive pursuit, made available only recently due to technologic advancements. Such cell-based assays were developed by researchers at Institut Pasteur Korea and have met widespread use [15, 16, 17, 18]. This assay employed the use of wild type parasites of T. cruzi infecting a non-modified cell line and the imaging of the resulting infection (in the amastigote stage) by high-content analysis (HCA) microscopy. Furthermore, the assay was developed in the 384-well format, allowing a high throughput testing of compounds. Preliminary cell toxicity is concomitantly determined by quantifying the ratio of host cell nuclei, a clear advantage since it reduces the need of an independent assay to assess this parameter. Using this screening assay, the authors of this chapter have also screened a library of 4000 kinase/phosphatase-like inhibitors that allowed the identification of 11 compounds with strong anti-parasitic activity and selectivity, suitable for follow up hit-to-lead optimization (unpublished results). In addition, a complementary assay developed for phenotypic profiling also allowed the identification of several compounds that interfered with the development and intracellular differentiation of T. cruzi. Compounds that hindered the differentiation from trypomastigotes to amastigote and the replication of amastigotes inside host cells are among the examples of “phenotypic” hits discovered (unpublished results). Due to the complex genetics and still many unknown aspects of T. cruzi biology, these types of compounds have the potential to constitute important chemical genomic tools that may help answering fundamental questions like: what triggers stage differentiation and what are the pathways involved? What factors are responsible for parasite persistence? How are amastigotes kept dormant for years to decades in host cells, hidden from the immune system? Due to the nature of the chemical library, it is likely that the compounds target kinases, of which the T. cruzi genome has 190 annotated potential members [19]. T. cruzi and other Trypanosomatids have a relatively big kinome when compared with other parasites that undergo several stage differentiations and contact with distinct environments, like Plasmodium spp. [20, 21]. One hypothesis is that while in metazoa and yeast the ultimate targets of many signaling cascades are transcription factors, which then trigger the expression of new sets of genes, Trypanosomatids have constitutive transcription of a majority of genes in large polycistronic units, hinting at a greater role of post-translational modifications (PTMs) like phosphorylation and acetylation.

Advertisement

2. Sirtuins: family and functions

Post-Translational Modifications (PTMs) represent one of the major mechanisms in regulating protein function in all life forms. Through phosphorylation, acetylation, methylation, glycosylation and ubiquitination, cells greatly extend the possibilities beyond the coding genome [22]. PTMs can change the enzymatic activity of a protein, change its subcellular localization, interfere with protein complexes assembly, increase or decrease its stability and induce interactions with DNA and RNA [22, 23].

Discovered half a century ago and largely ignored for the following years, lysine acetylation has re-emerged in the last two decades as a highly important PTM [24, 25]. Initial studies had focused in the role of lysine acetylation in the regulation of chromatin structure and gene expression, but with the advances in proteomic approaches, it was possible to begin to explore the function of lysine acetylation in non-histone proteins [24, 25].

Studies based on proteomic analysis to describe the lysine-acetylated proteins repertoire of an organism, called acetylome, have shown the presence of lysine acetylation in proteins from different cellular compartments and involved in different biological processes in several organisms [26]. Because of that, lysine acetylation has been placed by some authors in the same level of biological relevance as phosphorylation [24, 25]. In fact, studies of the acetylome of mammalian cells revealed acetylation sites in 1750 different proteins [27], a number close to the about 2000 proteins found to be phosphorylated [28].

The “acetyl code” is maintained by three different protein types: the “writers”, lysine acetyltransferases (KATs) that add acetyl groups to proteins, the “erasers”, lysine deacetylase (KDACs) that remove acetyl groups, and “readers”, proteins that specifically recognize and bind acetyl-lysine groups [26].

KDACs in particular have been the focus of great interest by the scientific community due to their many roles in cell function and disease. KDACs are interchangeably called histone deacetylases (HDACs), because the first discovered reactions catalyzed by these proteins were the removal of acetyl groups in histone tails [24, 29].

HDACs are separated into four different classes based upon sequence homology (class I, II, III and IV) and two different families: the histone deacetylase family and the sirtuin family, the latter being all class III HDACs. While the first family has a limited set of molecular targets, mainly composed of histones, sirtuins have a variety of substrates ranging from metabolic enzymes to structural proteins, as well as histones [30, 31, 32]. The sirtuin family seems to be ubiquitous throughout all kingdoms of life. The number of genes coding for sirtuins within an organism ranges from as little as one in bacteria, to seven in vertebrates [33]. The sirtuin family is further classified in 5 subclasses (I, II, III, IV and V) [34].

The most common reaction catalyzed by sirtuins is deacetylation. This reaction is of upmost biological importance as there is a clear relation between the acetylation status of several proteins and their cellular functions [35, 36, 37, 38]. The deacetylase reaction requires (nicotinamide adenine dinucleotide) NAD+, an acetylated lysine residue and produces deacetylated lysine, nicotinamide and 2′-O-acetyl-ADP-ribose (OAADPR) [39]. Studies on the kinetics and biochemical properties of the enzymes revealed binding to the acetyl-lysine substrate prior to NAD+. This is followed by nicotinamide cleavage from NAD+, that is the first product released, followed by deacetylated lysine and OAADPR [40] (Figure 2A). All sirtuins are strictly NAD+ dependent, a distinct characteristic that distinguishes them from other deacetylases. In fact, SIRT6 is a sirtuin capable of tightly binding to NAD+ without the requirement of an acetylated substrate, indicating that it may function as a NAD+ sensor [41].

Figure 2.

Sirtuin deacetylation and ADP-ribosylation mechanisms. (A) Sirtuins carry out protein deacetylation by removing acyl chains from protein lysine residues. This activity requires the cofactor nicotinamide adenine dinucleotide (NAD+), such that nicotinamide is released with concomitant production of O-acetyl-ADP-ribose (O-AADPR). There is a sirtuin-independent equilibrium between 2′ and 3′-O-AADPR isomers. The unspecified stereochemistry at the Cl′ position of O-AADPR reflects the fast epimerization observed in solution between α and β-anomers. (B) A number of sirtuins also exhibit ADP-ribosylating functionality. This reaction is also NAD+-dependent, as the cofactor acts as the source of ADP-ribose. ‘X’ represents the nucleophilic side-chain of a protein residue.

Besides being an endogenous product of the deacetylation reaction, nicotinamide is also a well-known inhibitor of sirtuins. Nicotinamide is an amide of nicotinic acid (vitamin B3) and is part of common enzyme co-factors like NAD+ and NADP (nicotinamide adenine dinucleotide phosphate) [42]. Intracellular physiological levels of nicotinamide in some mammalian cells seem to be in the range similar to the IC50’s of some sirtuins reinforcing the hypothesis that some sirtuins may act as NAD+ and nicotinamide sensors [43, 44].

OAADPR is another product of the deacetylation reaction [45] (Figure 2A). Early studies characterizing this molecule found that quantitative microinjection into starfish oocytes led to a blockage of oocyte maturation, indicating for the first time that OAADPR can evoke a biological activity [46]. There is now mounting evidence that OAADPR can elicit downstream responses that might synergize or antagonize the biological functions of sirtuin genes. So far, OAADPR has demonstrated to be related with functions in gene silencing, ion channel modulation and cell redox state maintenance [45].

Another reaction catalyzed by sirtuins is ADP-ribosylation. Although sirtuins were firstly described as ADP-ribosyltransferases (Figure 2B), their deacetylase activity has quickly overshadowed this activity, and as a consequence the biological processes associated to this reaction remains poorly understood [47]. In truth, sirtuins may have just deacetylase activity, both, or be mostly ADP-ribosyltransferases. An example is SIRT4 that is an efficient in vitro ADP-ribosyltransferase of histones that only recently was discovered to possess deacetylase activity [48, 49]. There is an active debate on whether ADP-ribosylation is in fact a biologically relevant function of sirtuins, or just an irrelevant side reaction/non-enzymatic artifact [50]. Nevertheless, some of the players in the dynamics of intracellular ADP-ribosylation have only been recently identified and it is becoming apparent that this PTM might be relevant for the modulation of important cell processes and signaling pathways like signal transduction mechanisms, transcription and DNA repair [51]. The ADP-ribose hydrolysis in Trypanosomatids has recently been studied in both Trypanosoma brucei and Trypanosoma cruzi and found to be mediated by a macrodomain with a conserved catalytic site [52].

Although (de)acetylation is the most common PTM, (de)acylation of other groups can be targeted, like succinyl and malonyl groups. SIRT5 and SIRT6 are some examples of proteins that perform deacylations of lysines other than acetyl groups, and their activities are important regulators of cell functions [53, 54, 55].

There is a wealth of information regarding the structural features of sirtuins. To date, some 83 structures of sirtuins are available in the protein databank, many of them co-crystalized with natural ligands or inhibitors. Though the available structures range from bacterial to mammalian sirtuins, the majority of the structures originate from the human genome.

Although the sequence homology can vary significantly between sirtuins, especially between prokaryotic and eukaryotic proteins, there is a conserved catalytic core of about 250 amino acids common to all members in the family [56]. The structure similarity of the Plasmodium falciparum PfSir2A with the mammalian SIRT5, despite a sequence homology of just 33% is a clear illustration [56]. This core contains a Rossman fold domain that is a NAD+ binding site, and a Zn2+ binding domain containing four highly conserved cysteine residues arranged in the motif (CX2CX20CX2). The catalytic site is located inside a hydrophobic channel formed at the interface of these two domains [56].

Whereas in HDACs from class I, II and IV, Zn2+ is an active participant in catalysis by producing free acetate and deacetylated lysine, in sirtuins it does not participate in reaction. However, the metal is essential for structural integrity, as was elucidated by the reversible loss of activity in a P. falciparum sirtuin where the zinc ion was removed [57]. Interestingly, an exception to the conservation of the CX2CX20CX2 motif is found on some sirtuins of Trypanosomatids, where one of the cysteines is not present [58]. However, deacetylase activity does not seem to be affected [59]. The molecular mechanism of deacetylation is still not completely elucidated, but it is generally accepted that the first step in the reaction involves the nucleophilic addition of the acetyl oxygen to nicotinamide ribose by a mechanism of SN2 substitution to produce O-alkylamine intermediate and nicotinamide. Then the acetyl group is transferred to ADP-ribose to form O-acetyl-ADP-ribose and deacetylated lysine [60].

The founding member of the sirtuin family is Sir2 from the budding yeast and was initially identified as part of a protein complex necessary to silence the expression of the mating-type-loci [61, 62]. Subsequently, it was also implicated in transcriptional silencing at telomere proximal sites [63] and ribosomal repeats at the ribosomal DNA (rDNA) locus [64, 65, 66]. Sir2 can be associated in distinct protein complexes that vary according to target site. For instance, at telomeres and the mating-type-loci, Sir2 forms a complex with two other homologs, Sir3 and Sir4 [63], while at rDNA locus Sir2 associates with Net1 and Cdc14 to form the regulator of nucleolar silencing and telophase exit—RENT complex [67, 68]. Yeast cells lacking Sir2 present a reduced lifespan that has been correlated with the accumulation of extrachromosomal ribosomal DNA circles originating from illegitimate recombination that are toxic to the cell and have been associated with aging [66, 69, 70].

In mammals, the nuclear SIRT1 is the most extensively studied member among the sirtuin family. The TATA box binding protein-associated factor RNA polymerase I subunit B (TAFI68) was the first substrate to be identified for SIRT1 in mouse cells. It is a transcription factor necessary for regulating the RNA polymerase I transcriptional complex, where it was shown that deacetylation inhibits transcriptional initiation in vitro [71]. Studies on p53, a non-histone substrate, demonstrated that acetylation activates the DNA-binding activity and target gene expression, also increasing its stability [72]. Consistent with this proposed SIRT1 inhibition of p53 function, SIRT1 knockout mice exhibit p53 hyper-acetylation and increased radiation-induced apoptosis, suggesting that SIRT1 can facilitate tumor growth by antagonism of p53 [73]. Still, the fact that SIRT1 can be found either overexpressed or underexpressed in different tumor types, and the finding that it can also function as a tumor suppressor [74] has hindered the clarification of its role in tumorigenesis. SIRT1 also plays an important role in metabolism, and its relation with caloric restriction and life-span extension has received much attention (reviewed in [75, 76, 77]). The beneficial effects of caloric restriction have been focused on the insulin-like growth factor-1 (IGF-1) and the target of rapamycin (TOR) pathways [78, 79], but increasing evidence suggest a role of SIRT1 in caloric restriction in mammals as well. For instance, SIRT1 expression was found to be elevated in models of caloric restriction, like fasting mice, low calorie diet in rat, or humans on a 25% caloric restriction diet [80, 81, 82]. On the other hand, mice lacking SIRT1 lost the life-span extension benefits of a 40% reduced calorie diet [83]. Despite many studies, the exact molecular mechanisms of SIRT1 in caloric restriction are still to be unraveled.

In vitro studies attribute a role to human SIRT2 in cell cycle regulation through the deacetylation of both tubulin and histone H4 [32, 84, 85]. In particular, it has been found that SIRT2 overexpressing cells were significantly delayed in cell cycle progression through mitosis [86]. Some links with age-related diseases have been described for SIRT2, such has neurodegenerative diseases [87, 88, 89], and different kinds of cancer. Mice lacking SIRT2 are prone to the appearance of tumors, an effect believed to be mediated by SIRT2 negative regulation of the anaphase-promoting complex [90]. It was demonstrated that SIRT2 expression is reduced in human gliomas, some of the most frequent malignant tumors in the brain [32, 91].

SIRT3 is the major mitochondrial deacetylase and studies with double knockout mice have revealed high levels of acetylation in protein targets [92, 93]. In addition, it was observed that these mice have impaired production of ATP [92]. When fasting or fed with a high-fat diet, the mice display atypical phenotypes that include cold intolerance and decreased ketone body formation [94, 95]. This strengthened the link with thermogenesis that had been previously demonstrated [96, 97]. In fact, SIRT3 expression is induced in mice in both white and brown adipose tissue upon caloric restriction and exposure of brown adipose tissue to cold temperatures [98]. In addition, SIRT3 also has a role in the deacetylation and activation of fatty acid β-oxidation, amino acid metabolism, electron transport chain and antioxidant defenses [99, 100].

SIRT4 was originally thought to be an unusual sirtuin due to the lack of deacetylase activity [101]. However, it was shown to ADP-ribosylate and down-regulate glutamate dehydrogenase production of ATP and has been implicated in insulin regulation of β-cells [48, 101]. Moreover, SIRT4 has recently been attributed a tumor suppressive function due to its involvement in DNA damage protection mediated by inhibition of mitochondrial glutamine metabolism, suggesting it might have therapeutic potential for treating glutamine-dependent cancers [102]. This mechanism is inhibited by mammalian target of rapamycin complex 1 (mTORC1) pathway [103]. SIRT4 also coordinates the balance between lipid synthesis and their catabolism by repressing malonyl-CoA decarboxylase, proving that it has, in fact, deacetylase activity [49].

SIRT5 is a NAD+-dependent protein lysine demalonylase and desuccinylase [104] and also has a deglutarylase activity [105]. It has a deacetylase activity [30], but has preference for acyl-carboxyl negatively charged groups [104, 105]. Some of its functions are related to glycolysis modulation [55]. The succinylome of mammalian cells has revealed many points of succinylation that are possible targets of SIRT5, mostly concentrated on mitochondrial metabolism [54]. SIRT5 also promotes urea cycle function via the regulation of carbamoyl-phosphate synthase [30, 35], and purine metabolism via urate oxidase [106]. Although a global protein hypersuccinylation and elevated serum ammonia during fasting were observed in SIRT5 knockout mouse model, the enzyme deficiency did not lead to any major metabolic abnormalities under either low or high fat diet conditions. These observations suggest that SIRT5 is likely dispensable for metabolic homeostasis under the basal conditions. It remains to be evaluated the role of SIRT5 in extreme conditions [107].

While most mammalian sirtuins have been implicated with metabolism, SIRT6 seems to be the only one with a direct link supporting a defined role in mammalian aging [108]. In fact, mice lacking SIRT6 gene develop a progeroid-like symptom with loss of subcutaneous fat, curved spine and lymphopenia. They develop normally for 2 weeks after birth, but then suffer from acute degeneration processes, ending up dying at 1 month of age [109]. At first considered to not possess deacetylase activity, but solely an ADP-ribosyltransferase activity [110], it was later found that SIRT6 removes both acetyl and long chain acyl groups from target molecules [53, 111]. It is localized in the nucleus, associated with chromatin, where it promotes the specific NAD+-dependent deacetylation of H3K9 and H3K56 [111, 112, 113]. SIRT6 is involved in genome protection by assuring correct telomere maintenance [111, 112], as well as DNA repair by double-strand break repair machinery [114, 115]. Like other sirtuins it also has a role in metabolism by influencing both glycolysis and gluconeogenesis [116, 117, 118] and lipid metabolism, by regulating triglyceride synthesis [108, 119]. Conditions like inflammation, heart disease and cancer all seem to be linked with SIRT6 function [120].

SIRT7 is the least studied sirtuin of all the mammalian sirtuins, but recent findings have established new functions and roles for this protein. It is a nucleolar sirtuin [121] and its localization is associated with the main process happening at this sub-nuclear structure, namely rDNA transcription [122]. SIRT7 does not possess a very strong deacetylate activity toward common synthetic and natural peptides [121], which is in agreement with the fact that SIRT7 depletion in mice did not change the global acetylation levels of either nucleus or nucleolus proteome [123], indicating that SIRT7 deacetylase activity is specific to a limited set of proteins. One example is specific deacetylation of H3K18 [31, 124] that underlies its role in chromatin remodeling. Also, SIRT7 has been found to be closely associated with B-WICH complex, a chromatin-remodeling complex [125]. It also has a role in protein synthesis [123, 126] and contributes to cell survival, namely by protecting against genomic insult [127, 128], hypoxia [129] and low glucose induced stress [130]. All the functions described characterize SIRT7 as a pro-survival protein. Indeed, it is currently considered to be an oncogene in all the cancer types studied so far [126, 131, 132].

Advertisement

3. Parasitic sirtuins

Various genome-sequencing projects demonstrated the presence of genes coding for sirtuins in most protozoan parasites of medical importance. An interesting finding was that depending on the protozoan parasite species the number of sirtuins varied (Table 1).

Table 1.

Sirtuin genes identified in genome-sequencing programs for parasitic protozoa.

Plasmodium spp. have been shown to contain two sirtuin orthologues, called Sir2A and Sir2B. Sir2A is the most extensively studied homolog, mainly located at the nucleus [133] although it can also shuttle to cytoplasm after posttranslational SUMOylation [134]. Sir2A has been characterized as a mediator of transcriptional silencing at the telomeric regions of chromosomes [133, 135]. The telomeres of P. falciparum are rich in gene families involved in antigenic variation such as the var. family of genes. These genes are responsible for the expression of parasite-derived P. falciparum erythrocyte membrane protein, PfEMP1, responsible for immune evasion in humans [136]. The family of var. genes is tightly regulated by sirtuins, with the expression of its members being mutually exclusive [137, 138]. The switch of active var. is controlled exclusively at the epigenetic level [137, 139].

PfSir2B is a larger homolog with a molecular weight more than four times the size of Sir2A and is involved in the transcriptional silencing of a complementary subset of var. genes with distinct promoter types [140].

Sirtuins from Leishmania spp. parasites were among the first to be identified in Trypanosomatids, when a complementary DNA (cDNA) isolated and sequenced from Leishmania major showed a high homology with yeast Sir2 [141]. Antibodies raised against this LmSir2 later showed to be reactive against different life cycle stages of L. major, but also to Leishmania amazonensis and even to the serum of a patient infected with Leishmania infantum [142]. Furthermore, the protein was found to be among the secreted material of L. major [142].

Overexpression of the Sir2 protein in L. infantum, sharing 93% homology to the L. major protein, led to an increased survival of amastigotes under axenic conditions [143]. Also, when the overexpression was performed in mammalian fibroblasts, host cells became more permissive to infection by Leishmania infection in comparison with wild type cells, hinting at a modulation of host cell by the parasite [144]. Genetic knockouts in L. infantum of the Sir2 related protein 1 (Sir2rp1) gene also highlighted the importance of this protein in the parasite. While single knockouts were readily obtainable, double deletion of the alleles was only possible after the rescue by an ectopical copy of the gene, suggesting an essential role for parasite survival [145]. When single-knockouts of L. infantum Sir2rp1 were used to infect a macrophage cell line, in vitro, it was noted that although they had the same invasive capacity than wild-type parasites, they showed a hindered replication rate leading to diminished infections over-time. Furthermore, the mutant parasite also failed to establish an infection in an in vivo mouse model of Leishmaniasis [145]. Cellular and biochemical studies later established LiSir2rp1 has NAD+-dependent deacetylase with ADP-ribosylation activity that co-localized to the cytoskeleton and potentially interacted with tubulin as well as with HSP83, an orthologue of mammalian HSP90 [146, 147]. The association with cytoskeleton is a characteristic feature of both SIRT2 and HDAC6 in mammalian cells [32, 148]. In addition, an orthologous from L. amazonensis, LaSir2rp1 was found to be a glycosylated protein, but whether this is the case for other species remains to be seen [149]. Although the Sir2 related protein 1 has received much attention, no studies have been made for the other two proteins codified by the Leishmania species; Sir2 related protein 2 and Sir2 related protein 3.

Sir2rp1 from L. donovani has also been implicated in the resistance of amphotericin B, a reference drug in the treatment of visceral Leishmaniasis. When clinical isolates were targeted for gene knockout of the protein, parasites showed a lower level of multi-drug resistant transporter MDR1, lower drug efflux, increased ROS production and increased sensitivity to amphotericin B [150]. On the contrary, overexpression led to a resistant phenotype, thereby suggesting Sir2 as a new resistant marker for visceral Leishmaniasis [150]. Comparative transcriptomic analysis also implicates Sir2 in resistance to miltefosine, another drug used to treat the disease [151]. In addition to its potential as a novel drug target, Sir2rp1 from Leishmania spp. has also been suggested as a vaccine [152, 153].

Recently, LiSir2rp2 and LiSir2rp3, the others Leishmania sirtuins, were characterized as mitochondrial proteins, and while LiSir2rp3 was demonstrated to not be essential (Table 1), attempts to generate LiSir2rp2 knockout cells failed. LiSir2rp2 was implicated in parasite proliferation depending on NAD+ availability [154].

Trypanosoma brucei, like Leishmania spp., has 3 sirtuins annotated in its genome (Table 1). The first enzyme to be characterized in the parasite was TbSir2rp1 [155]. The enzyme is localized mainly in the nucleus in association with chromosomes. The protein was shown to possess both deacetylase activity toward endogenous histones while being also able to ADP-ribosylate calf thymus histones and, to a lesser extent, bovine serum albumin (BSA). Up to that time, no ADP-ribosylation had been detected in common members of the sirtuin family like yeast Sir2 or HST2, hence TbSir2rp1 was one of the first enzymes to exhibit this dual activity [156, 157, 158, 159]. Furthermore, mutation of a catalytic histidine essential for deacetylase activity also affected ADP-ribosylation activity, suggesting that the two activities were occurring simultaneously. Because of the increased or decreased resistance to DNA damage caused by the alkylating drug methyl methanesulfonate (MMS) in overexpressing or RNAi-induced knockdown T. brucei cell lines, respectively, TbSir2rp1 was also considered to have a role in DNA repair in this organism [155].

A later study performed with bloodstream forms (as opposed to insect stage forms in the previous works) characterized TbSir2rp1 and also the other two sirtuins, TbSir2rp2 and TbSir2rp3 that both with mitochondrial localization [160]. TbSir2rp1 was found in the nucleus, but when overexpressed to high levels in T. brucei cells, it localizes to the cytoplasm, with toxic effects to the parasite [160]. Besides, gene knockouts for the three proteins caused no growth in parasites maintained in standard conditions [160]. TbSir2rp1 mutants, however, did show an increased sensitivity to MMS damage, confirming the previous results performed with RNA interference (RNAi). The particular localization of TbSir2rp1 led to the investigation of Sir2 mediated telomere gene silencing like the one that occurs in yeast and Plasmodium spp., as discussed earlier [161, 162]. Although TbSir2rp1 was found to have a role in telomere DNA repair and telomere silence, it was not required for antigenic variation [160] as described for Plasmodium Sir2 [133, 135].

TbSir2rp1 has also been studied as a model sirtuin, with both deacetylase and ADP-ribosylation activity. Biochemical experiments revealed that ADP-ribosylation is 5-fold less active than the deacetylation reaction, and occurs only in the presence of an acetylated substrate by two distinct biochemical mechanisms [163]. Another research group demonstrated that ADP-ribosylation can also occur in arginine, independent of the presence of an acetylated substrate, as supported by mass spectrometry and molecular dynamics simulations [164].

Additional information about parasitic sirtuins can be found in a recent review by Hailu et al. [165].

Advertisement

4. Trypanosoma cruzi sirtuins

Although the draft of the T. cruzi genome has been published a decade ago [19], it was not until recently that the first experimental studies involving the sirtuins of this parasite have been published. Unlike Leishmania spp. and T. brucei that possess three Sir2-like proteins, T. cruzi only has two coding sequences annotated in its genome, TcSir2rp1 and TcSir2rp3 (Table 1).

One study employing parasites overexpressing TcSir2rp1 and TcSir2rp3 by a tetracycline inducible vector characterized some of the features of both proteins [58]. Localization studies employing both wild type parasites and polyclonal sera raised against both proteins, as well as localization of tagged copies in the overexpression mutants with monoclonal antibodies, attributed a cytoplasmic localization to TcSir2rp1 and a mitochondrial to TcSir2rp3 [58]. Both of the proteins’ levels are regulated throughout the life cycle of the parasite, with a significant decrease in both at the trypomastigote stage [58]. Overexpression of TcSir2rp1 was responsible for higher metacyclogenesis and higher infectivity of Vero cells. Since metacyclogenesis occurs under nutrient deprivation, it is hypothesized that TcSir2rp1 may function like sirtuins from other organisms that respond to starvation [101, 166, 167]. On the other hand, overexpression of TcSir2rp3 led to a decrease in epimastigote replication time, lower infectivity in Vero cells, increased amastigote replication and normal metacyclogenesis [58]. Due to the oxidizing environment in which amastigotes replicate, it has been suggested that TcSir2rp3 performs protecting functions against oxidative stress like SIRT3 [36]. Both of the overexpressing cell lines reduced the levels of acetylation for particular proteins, as well as protected against the effect of specific sirtuin inhibitors [58].

Moretti and colleagues also independently characterized both of the sirtuins in a simultaneous study [59]. In their study, they show that salermide, a sirtuin inhibitor analogue of sirtinol found to be a strong anticancer molecule, is active against both in vitro cultures of epimastigotes, and against an in vivo model of infection by T. cruzi, albeit at moderate levels [59, 168]. Salermide was also found to be a strong inhibitor of TcSir2rp3 recombinant protein [59]. The authors report the same localization for both proteins, as well as the interference in epimastigote growth, metacyclogenesis, infectivity of host cells and amastigote replication in lines overexpressing the sirtuins. Differently from Ritagliati work [58], in their studies, the overexpression of the cytosolic TcSir2rp1 caused a decreased in the epimastigotes proliferation while TcSir2pr3 increased the growth rate. These differences might be due to the amount of overexpression achieved and parasite strains used [59].

Advertisement

5. Potential of Trypanosoma cruzi sirtuins as targets for infection control

A strategy that has been traditionally employed in Chagas disease drug discovery is the target-based approach. One such molecular target that has gained increasing interest as a potential drug targets against parasitic diseases is that of sirtuins [169, 170]. The hypothesis arose by the time that it was demonstrated that sirtuins are life-span regulators in organisms like yeast, flies and worms [70, 171, 172]. Therefore, many groups promptly investigated whether sirtuin orthologues present in parasites could have important functions that could be exploited for novel therapeutic applications.

One important aspect for the viability of targeting sirtuins in parasites is the homology between the protein of interest and other proteins present in the host organisms. Although sirtuins are conserved through evolution [33], significant difference at the sequence level can be found between Trypanosomatid and human homologs. For instance, T. cruzi Sir2rp1 shares only 33% identity with mammalian SIRT2, its closest homolog (Multi-way protein alignment, BLOSUM 62) [173].

Another argument that has led to the consideration of T. cruzi sirtuins as a drug target is that this family of proteins is considered to possess structural properties adequate to inhibition by small-molecule compounds. In particular, the catalytic site is located inside a hydrophobic channel formed at the interface between the two constituting domains, the Rossman fold containing the NAD+ binding domain and the Zn2+ ion binding domain [56]. Catalytic pockets buried inside the protein are considered an essential feature for target druggability [174].

One last fact that prompted the evaluation of TcSir2rp1 as a drug target was the previous evidence that a class of experimental compounds preferentially inhibited LiSir2rp1 over the human homolog SIRT1 [175]. The possibility to synthetize selective sirtuin inhibitors has been successfully achieved for human homologs, based upon structural knowledge of the catalytic site as has been demonstrated for human SIRT2 [176, 177].

Enzymatic inhibition by small molecule compounds is an essential step in the druggability assessment of novel therapeutic targets [174]. Biochemical studies performed by our research groups evaluated the effect of nicotinamide, a classic non-competitive inhibitor of sirtuins in TcSir2rp1 and TcSir2rp3 (to be published). TcSir2rp1 was shown to be inhibited by nicotinamide, albeit at a relatively high IC50 when compared with other sirtuins (4-fold higher for hSIRT1 and 11-fold higher for LiSir2rp1) [175]. Different nicotinamide sensitivities are found among distinct sirtuins, and may explain the differences described [177]. Nicotinamide inhibits deacetylation by binding to a conserved C pocket present in sirtuins that participates in NAD+ binding and catalysis, where it promotes a base-exchange reaction at the expense of deacetylation [178]. A hypothesis for the high IC50 for nicotinamide in TcSir2rp1 could be related with structural characteristics of this conserved C pocket. Structural determination of TcSir2rp1 by X-ray crystallography currently ongoing in our group will certainly highlight these differences. Contrary to previous studies [179], we could not observe any antiparasitic activity of nicotinamide against T. cruzi amastigotes for up to a concentration of 2 mM. Several studies report the activity of nicotinamide against parasitic protozoa [180, 181, 182], but to our knowledge, none clearly establishes a relation between antiparasitic activity and sirtuin inhibition.

Other biochemical functions and protein interactions have been attributed to Sir2rp1 in related Trypanosomatids, and future experiments should shed light whether it applies to TcSir2rp1. One of the biochemical functions that has been characterized for both TbSir2rp1 and LiSir2rp1 is ADP-ribosylation. Both orthologous showed to ADP-ribosylate calf thymus histones and BSA [155, 183]. Later studies involving TbSir2rp1 demonstrated that this biochemical function is dependent upon acetylated histones, is coupled to the deacetylase activity of the sirtuin, but occurs at a much lower rate than the latter [163]. In fact, even though ADP-ribosylation has clear functions in both physiological and pathogenic situations when catalyzed by other ADP-ribosyltransferases [183, 184], the reaction catalyzed by sirtuins is currently challenged to be an unspecific side-reaction [159].

Like the human SIRT2, L. infantum Sir2rp1 was also found to be associated with tubulin [146], the major component of Trypanosomatids cytoskeleton formed by an array of subpellicular microtubules that span the whole cell of the parasite [185]. SIRT2 is a tubulin deacetylase that displays a higher affinity for tubulin than for histones [32], and has been found to be linked to regulation of mitotic progression [86], chromatin condensation [186] and cell migration [187]. TcSir2rp1 overexpression in T. cruzi was found to increase the deacetylation level of endogenous tubulin [58]. It is interesting to note that Sir2rp1 from T. cruzi is a cytoplasmic protein like LiSir2rp1 and not nuclear like TbSir2rp1. Since T. cruzi shares some characteristics with L. infantum like the amastigote intracellular stage, it should not be ruled out that Sir2rp1 may have functions in the cytoskeleton remodeling necessary for stage differentiation. Several proteins, sirtuins included, have demonstrated the ability to shuttle from the nucleus to the cytoplasm and vice-versa [188, 189]. SIRT2, the closest sirtuin homolog of mammalian cells, is actively exported to the cytoplasm during interphase, but is accumulated in the nucleus from prophase until cytokinesis where it co-localizes with important mitotic structures like centrosomes and the mitotic spindle [190]. Curiously, analysis of TcSir2rp1 by Wregex and cNLS Mapper, bio-computational tools that identify nuclear export signals (NES) and nuclear localization signals (NLS), respectively, indicate the presence of non-canonical NES/NLS in the sequence of this sirtuin [191, 192]. Whether TcSir2rp1 does shuttle to the nucleus during specific phases of T. cruzi life cycle, for instance to repair DNA damage like the T. brucei orthologue, remains to be reported.

The mitochondrial TcSir2rp3 was found less expressed in T. cruzi forms proliferating in mammalian cells. Its expression increased when the parasite transformed in trypomastigotes [59]. The fact that cells overexpressing only the active form, but not the inactive form of TcSir2rp3 showed an increased intracellular growth and failed to transform in to extracellular trypomastigotes [59], suggested that it could also be a drug target to control the infection, although further experiments to generate knockout cell lines need to be performed. We recently identified several compounds that prevented intracellular growth of T. cruzi some of them inhibiting specifically TcSir2rp1 or TcSir2rp3 which might indicate the requirement of both enzymes for the parasite (to be published).

Advertisement

6. Naphtalimide derivatives as potential drugs for Chagas disease control

Naphthalimides are a class of compounds that have generated intense interest as active molecules with potential to treat a range of conditions [193]. A naphthalene ring linked to an imide group that forms a third heterocycle composes the basic chemical scaffold of the naphthalimide derivatives. This moiety has a planar nature and is considered to be responsible for the pharmacological activities attributed to compounds derived from this structure, that can be as distinct as anticancer, antibacterial, antiprotozoal, antiviral, analgesic, and anesthetic [193]. Their potential as anticancer compounds has received particular attention, mostly because of their DNA intercalating properties and also to their reported activity as topoisomerase inhibitors [194, 195, 196]. Compounds like amonafide and bisnafide have been proposed as anticancer agents and have inclusively reached clinical trials in the past [197, 198]. Elinafide is another derivative with two naphthalimide moieties that has been evaluated in preclinical studies and demonstrated potential against various mouse xenograft models [199]. This last compound was in the origin of the synthesis of the first BNIPs that differed in the alkyl chain linking the naphthalimide and amine group, i.e. a propyl instead of an ethyl chain [200]. These derivatives showed potential activity against breast cancer MCF-7 cell line and actively bound DNA as demonstrated by thermal denaturation measurements, ethidium bromide displacement and DNA gel mobility [200]. Later derivatives that varied in the length of the chain linking the two amines of bisnaphthalimidopropyl groups were also evaluated against cancer cell lines and promastigotes of the parasite L. infantum [201]. While screening for enzymatic inhibitors of the recently characterized LiSir2rp1, BNIPs were identified as inhibitors of its deacetylase activity [175]. Furthermore, they were active against intracellular amastigotes, the clinically relevant stage of the parasite present in humans, at concentrations in the single micromolar range [175].

Our groups’ previous results demonstrating activity toward L. infantum led to the testing of BNIPs as inhibitors of the related Trypanosomatid T. brucei and its Sir2rp1 orthologue, TbSir2rp1 [202]. BNIPs revealed to be very potent inhibitors of in vitro parasite growth, with one of the compounds, BNIPDabut (Figure 1) with an EC50 in the range of the reference drug pentamidine. However, when tested against the TbSir2rp1 recombinant enzyme, BNIPDabut had an IC50 more than 104 times superior to the EC50 against the whole cell parasite, indicating that Sir2rp1 inhibition is probably not a major mechanism of action for the compound. Whether BNIPDabut inhibits other T. brucei sirtuin enzymes remains to be elucidated. It should be noted that RNAi and gene knockout experiments of the three sirtuins did not led to a deleterious effect, and unlike LiSir2rp1, there is no indication of that these proteins may be essential [155, 160]. The optimal in vitro properties of BNIPDabut led to the testing with an in vivo model of trypanosomiasis by bioluminescence imaging. Although BNIPDabut maintained a strong trypanocidal activity in vivo, as assessed by the decrease in bioluminescent signals to levels similar to those of the reference drug control pentamidine, it was not sufficient for infection clearance, as animals’ parasitaemia relapsed shortly after treatment interruption. Nevertheless, BNIPDabut should constitute a scaffold for further consideration in HAT drug discovery [202].

In a different study, BNIP derivatives were also tested against TcSir2rp1 and in vitro cultures of T. cruzi amastigotes (unpublished results). BNIPs demonstrated to inhibit the deacetylase activity of this enzyme, with BNIPSpd (Figure 1) as the most potent compound, showing a dose-dependent effect on inhibition. BNIPSpd also proved to be active and selective against amastigotes of T. cruzi in a high content screening (HCS) assay – an assay that takes advantage of computer-assisted image processing of hundreds of microscopic images originating in automated microscopes able to simultaneously capture multiple conditions or drug compounds. In this work, a new set of derivatives was also synthetized in order to improve both solubility and binding to cellular targets, mostly by including cyclic structures and heteroatoms in the carbon chain linking the two naphthalimide groups. The most active compounds were BCNIPP, also a TcSir2rp1 inhibitor, and trans BNIP-1,4-Dacyhex, a derivative of BNIPDabut that weakly inhibited the enzyme. In turn, BNIPDabut had some inhibitory activity on T. cruzi amastigotes, with low selectivity, but also did not inhibit TcSir2rp1 at 10 μM. The activity of BNIPSpd in a mice model of Chagas disease using bioluminescent parasites was also determined and found to be absent at the doses tested. An explanation might be the poor pharmacokinetic profile of the compounds, which fails to ever achieve at least the in vitro IC50 against T. cruzi amastigotes.

Altogether, our data indicate that BNIP derivatives may not be acting only by a mechanism of Sir2rp1 inhibition, with other targets contributing to the activity detected. BNIPs were originally designed and developed as anti-cancer agents [203, 204] through DNA intercalation. This property might explain some of the cytotoxic effects verified against host cells, but may also be an important mechanism of activity toward the parasite, especially since trypanosomes are highly susceptible to intercalating agents [205].

Confirmation of the mechanism of action can be undertaken by appropriate target deconvolution experiments [206]. The most common type of such experiments are biochemical methods that employ some variation of biochemical affinity purification, where the compounds are immobilized in a column, and allowed to interact with protein extracts, preferably previously fractionated. After stringent washing steps, the bound proteins are eluted and identified. Such strategy has been employed in the identification of small molecule activators of cryptochrome of mammalian cells [207]. A disadvantage is that there is a bias toward high affinity ligands, and when the target is relatively less abundant or has less affinity, important targets may not be detected. Furthermore, the washing steps may eliminate protein complexes that may be important for appropriate drug activity.

Genetic methods can also be valuable for target deconvolution. Gene knockouts and RNAi screens can be used to try to phenotype a compound’s effect [208]. Furthermore, if the mutant is hypersensitive to the compound in question, the evidence that the protein could be the target for the compounds would be strengthened. The validation of trypanothione synthetase and N-myristoyltransferase as drug targets against trypanosomes are examples where the differential sensitivity of an inhibitor in wild type, overexpression, and knockout mutants is clearly illustrated [209, 210]. An additional genetic strategy is based on the generation of resistant cell lines by culturing the parasites in increasing sub-lethal drug concentrations that are posteriorly sequenced to find mutated genes related to the mechanism of action for the compound [208]. Genetic methods have recently been employed in the search of the mechanism of action of oxaboroles [211], a class of compounds in development for HAT, but also active against T. cruzi [212, 213].

Chemical genomics can also be applied to the discovery of novel drug targets, as exemplified by the recent characterization of cytochrome b from T. cruzi. This enzyme was demonstrated to be selectively targeted in relation to the human homolog by a hit compound coming from a phenotypic screening [214].

Advertisement

7. Perspective

Here we highlighted the potential of sirtuins, particularly in the T. cruzi as possible targets of drug development for Chagas disease for the following reasons:

  1. The parasite contains only two sirtuins, instead of three present in Leishmania spp. and T. brucei species, and seven in the human host, which facilitates a more precise design and avoid redundant effects.

  2. Several experiments demonstrate the requirement of these enzymes for growth and survival, particularly in intracellular forms of the parasite, which are clinically relevant.

  3. There are several compounds already designed to be selective sirtuin inhibitors that could be modified to provide increased specificity and selectivity to the parasite enzyme. Some of these compounds have already defined druggability and specific derivatives can be repurposed more easily.

  4. Both sirtuins were produced to perform enzymatic screenings and crystals of TcSir2rp1 have been obtained to be the basis of further medicinal chemistry.

Finally, several lead compounds were identified for T. cruzi sirtuin, which can provide the basis for future development. It is also important to note that these possible inhibitors could act synergistically with drugs already in use for treatment, as a novel combinational therapy, opening new avenues to eliminate Chagas disease.

Advertisement

Acknowledgments

This article is a result of the project NORTE-01-0145-FEDER-000012, supported by Norte Portugal Regional Operational Programme (NORTE 2020), under the PORTUGAL 2020 Partnership Agreement, through the European Regional Development Fund (ERDF)”.

References

  1. 1. WHO. Chagas disease in latin america: An epidemiological update based on 2010 estimates. Weekly Epidemiological Record. 2015;90(6):33-43
  2. 2. Walsh JA. Estimating the burden of illness in the tropics. In: Warren KS, Mahmoud AAF, editors. Tropical and Geographical Medicine. McGraw Hill Education; 1984. pp. 1073-1085
  3. 3. Knowles LL, Carstens BC. Delimiting species without monophyletic gene trees. Systematic Biology. 2007;56(6):887-895
  4. 4. Deane LM. Animal reservoirs of Trypanosoma cruzi in Brazil. Revista Brasileira de Malariologia e Doenças Tropicais. 1964;16:27-48
  5. 5. Barreto M. Reservatórios do Trypanosoma cruzi nas américas. Revista Brasileira de Malariologia e Doenças Tropicais. 1964;16:527-552
  6. 6. Tarleton RL. Parasite persistence in the aetiology of Chagas disease. International Journal for Parasitology. 2001;31(5–6):550-554. DOI: 10.1016/S0020-7519(01)00158-8
  7. 7. Tarleton RL. Chagas disease: A role for autoimmunity? Trends in Parasitology. 2003;19(10):447-451
  8. 8. Wainwright M. Dyes, trypanosomiasis and DNA: A historical and critical review. Biotechnic & Histochemistry. 2010;85(6):341-354. DOI: 10.3109/10520290903297528
  9. 9. Docampo R, Moreno SN, Gadelha FR, de Souza W, Cruz FS. Prevention of Chagas’ disease resulting from blood transfusion by treatment of blood: Toxicity and mode of action of gentian violet. Biomedical and Environmental Sciences. 1988;1(4):406-413
  10. 10. Kennedy PG. The continuing problem of human african trypanosomiasis (sleeping sickness). Annals of Neurology. 2008;64(2):116-126. DOI: 10.1002/ana.21429
  11. 11. Gaspar L, Moraes C, Freitas-Junior L, Ferrari S, Costantino L, Costi M, et al. Current and future chemotherapy for Chagas disease. Current Medicinal Chemistry. 2015;22(37):4293-4312. DOI: 10.2174/0929867322666151015120804
  12. 12. Butera JA. Phenotypic screening as a strategic component of drug discovery programs targeting novel antiparasitic and antimycobacterial agents: An editorial. Journal of Medicinal Chemistry. 2013;56(20):7715-7718. DOI: 10.1021/jm400443k
  13. 13. Buckner FS, Verlinde CL, La Flamme AC, Van Voorhis WC. Efficient technique for screening drugs for activity against Trypanosoma cruzi using parasites expressing beta-galactosidase. Antimicrobial Agents and Chemotherapy. 1996;40(11):2592-2597
  14. 14. Canavaci AM, Bustamante JM, Padilla AM, Perez Brandan CM, Simpson LJ, Xu D, et al. In vitro and in vivo high-throughput assays for the testing of anti-Trypanosoma cruzi compounds. PLoS Neglected Tropical Diseases. 2010;4(7):e740. DOI: 10.1371/journal.pntd.0000740
  15. 15. Moon S, Siqueira-Neto JL, Moraes CB, Yang G, Kang M, Freitas-Junior LH, et al. An image-based algorithm for precise and accurate high throughput assessment of drug activity against the human parasite Trypanosoma cruzi. PLoS One. 2014;9(2):e87188. DOI: 10.1371/journal.pone.0087188
  16. 16. Moraes CB, White KL, Braillard S, Perez C, Goo J, Gaspar L, et al. Enantiomers of nifurtimox do not exhibit stereoselective anti-Trypanosoma cruzi activity, toxicity, or pharmacokinetic properties. Antimicrobial Agents and Chemotherapy. 2015;59(6):3645-3647. DOI: 10.1128/aac.05139-14
  17. 17. Balfour MN, Franco CH, Moraes CB, Freitas-Junior LH, Stefani HA. Synthesis and trypanocidal activity of a library of 4-substituted 2-(1h-pyrrolo[3,2-c]pyridin-2-yl)propan-2-ols. European Journal of Medicinal Chemistry. 2017;128:202-212. DOI: 10.1016/j.ejmech.2017.01.040
  18. 18. Moraes CB, Giardini MA, Kim H, Franco CH, Araujo-Junior AM, Schenkman S, et al. Nitroheterocyclic compounds are more efficacious than cyp51 inhibitors against Trypanosoma cruzi: Implications for Chagas disease drug discovery and development. Scientific Reports. 2014;4:4703-4714. DOI: 10.1038/srep04703
  19. 19. El-Sayed NM, Myler PJ, Bartholomeu DC, Nilsson D, Aggarwal G, Tran AN, et al. The genome sequence of Trypanosoma cruzi, etiologic agent of Chagas disease. Science. 2005;309(5733):409-415. DOI: 10.1126/science.1112631
  20. 20. Naula C, Parsons M, Mottram JC. Protein kinases as drug targets in trypanosomes and Leishmania. Biochimica et Biophysica Acta. 2005;1754(1–2):151-159. DOI: 10.1016/j.bbapap.2005.08.018
  21. 21. Ward P, Equinet L, Packer J, Doerig C. Protein kinases of the human malaria parasite Plasmodium falciparum: The kinome of a divergent eukaryote. BMC Genomics. 2004;5:79-98. DOI: 10.1186/1471-2164-5-79
  22. 22. Walsh CT, Garneau-Tsodikova S, Gatto GJ Jr. Protein posttranslational modifications: The chemistry of proteome diversifications. Angewandte Chemie (International Ed. in English). 2005;44(45):7342-7372. DOI: 10.1002/anie.200501023
  23. 23. Lothrop AP, Torres MP, Fuchs SM. Deciphering post-translational modification codes. FEBS Letters. 2013;587(8):1247-1257. DOI: 10.1016/j.febslet.2013.01.047
  24. 24. Verdin E, Ott M. 50 years of protein acetylation: From gene regulation to epigenetics, metabolism and beyond. Nature Reviews. Molecular Cell Biology. 2015;16(4):258-264. DOI: 10.1038/nrm3931
  25. 25. Kouzarides T. Acetylation: A regulatory modification to rival phosphorylation? The EMBO Journal. 2000;19(6):1176-1179. DOI: 10.1093/emboj/19.6.1176
  26. 26. Choudhary C, Weinert BT, Nishida Y, Verdin E, Mann M. The growing landscape of lysine acetylation links metabolism and cell signalling. Nature Reviews. Molecular Cell Biology. 2014;15(8):536-550. DOI: 10.1038/nrm3841
  27. 27. Choudhary C, Kumar C, Gnad F, Nielsen ML, Rehman M, Walther TC, et al. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science. 2009;325(5942):834-840. DOI: 10.1126/science.1175371
  28. 28. Olsen JV, Blagoev B, Gnad F, Macek B, Kumar C, Mortensen P, et al. Global, in vivo, and site-specific phosphorylation dynamics in signaling networks. Cell. 2006;127(3):635-648. DOI: 10.1016/j.cell.2006.09.026
  29. 29. Inoue A, Fujimoto D. Enzymatic deacetylation of histone. Biochemical and Biophysical Research Communications. 1969;36(1):146-150
  30. 30. Nakagawa T, Lomb DJ, Haigis MC, Guarente L. SIRT5 deacetylates carbamoyl phosphate synthetase 1 and regulates the urea cycle. Cell. 2009;137(3):560-570. DOI: 10.1016/j.cell.2009.02.026
  31. 31. Barber MF, Michishita-Kioi E, Xi Y, Tasselli L, Kioi M, Moqtaderi Z, et al. SIRT7 links h3k18 deacetylation to maintenance of oncogenic transformation. Nature. 2012;487(7405):114-118. DOI: 10.1038/nature11043
  32. 32. North BJ, Marshall BL, Borra MT, Denu JM, Verdin E. The human Sir2 ortholog, SIRT2, is an NAD+-dependent tubulin deacetylase. Molecular Cell. 2003;11(2):437-444. DOI: 10.1016/s1097-2765(03)00038-8
  33. 33. Frye R. Evolution of sirtuins from archaea to vertebrates. In: Verdin E, editor. Histone Deacetylases. Humana Press; 2006. pp. 183-202. DOI: 10.1385/1-59745-024-3:183
  34. 34. Frye RA. Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochemical and Biophysical Research Communications. 2000;273(2):793-798. DOI: 10.1006/bbrc.2000.3000
  35. 35. Ogura M, Nakamura Y, Tanaka D, Zhuang X, Fujita Y, Obara A, et al. Overexpression of SIRT5 confirms its involvement in deacetylation and activation of carbamoyl phosphate synthetase 1. Biochemical and Biophysical Research Communications. 2010;393(1):73-78. DOI: 10.1016/j.bbrc.2010.01.081
  36. 36. Bause AS, Haigis MC. SIRT3 regulation of mitochondrial oxidative stress. Experimental Gerontology. 2013;48(7):634-639. DOI: 10.1016/j.exger.2012.08.007
  37. 37. Bharathi SS, Zhang Y, Mohsen AW, Uppala R, Balasubramani M, Schreiber E, et al. Sirtuin 3 (SIRT3) protein regulates long-chain acyl-coa dehydrogenase by deacetylating conserved lysines near the active site. The Journal of Biological Chemistry. 2013;288(47):33837-33847. DOI: 10.1074/jbc.M113.510354
  38. 38. Sauve AA. Sirtuin chemical mechanisms. Biochimica et Biophysica Acta. 2010;1804(8):1591-1603. DOI: 10.1016/j.bbapap.2010.01.021
  39. 39. Jackson MD, Denu JM. 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. The Journal of Biological Chemistry. 2002;277(21):18535-18544. DOI: 10.1074/jbc.M200671200
  40. 40. Borra MT, Langer MR, Slama JT, Denu JM. Substrate specificity and kinetic mechanism of the Sir2 family of NAD+-dependent histone/protein deacetylases. Biochemistry. 2004;43(30):9877-9887. DOI: 10.1021/bi049592e
  41. 41. Pan PW, Feldman JL, Devries MK, Dong A, Edwards AM, Denu JM. Structure and biochemical functions of SIRT6. The Journal of Biological Chemistry. 2011;286(16):14575-14587. DOI: 10.1074/jbc.M111.218990
  42. 42. Berger F, Ramirez-Hernandez MH, Ziegler M. The new life of a centenarian: Signalling functions of NAD(P). Trends in Biochemical Sciences. 2004;29(3):111-118. DOI: 10.1016/j.tibs.2004.01.007
  43. 43. Smythe GA, Braga O, Brew BJ, Grant RS, Guillemin GJ, Kerr SJ, et al. Concurrent quantification of quinolinic, picolinic, and nicotinic acids using electron-capture negative-ion gas chromatography-mass spectrometry. Analytical Biochemistry. 2002;301(1):21-26. DOI: 10.1006/abio.2001.5490
  44. 44. Hagino Y, Lan SJ, Ng CY, Henderson LM. Metabolism of pyridinium precursors of pyridine nucleotides in perfused rat liver. The Journal of Biological Chemistry. 1968;243(19):4980-4986
  45. 45. Tong L, Denu JM. Function and metabolism of sirtuin metabolite o-acetyl-ADP-ribose. Biochimica et Biophysica Acta. 2010;1804(8):1617-1625. DOI: 10.1016/j.bbapap.2010.02.007
  46. 46. Borra MT, O’Neill FJ, Jackson MD, Marshall B, Verdin E, Foltz KR, et al. Conserved enzymatic production and biological effect of o-acetyl-ADP-ribose by silent information regulator 2-like NAD+-dependent deacetylases. The Journal of Biological Chemistry. 2002;277(15):12632-12641. DOI: 10.1074/jbc.M111830200
  47. 47. Frye RA. Characterization of five human cdnas with homology to the yeast Sir2 gene: Sir2-like proteins (sirtuins) metabolize NAD and may have protein ADP-ribosyltransferase activity. Biochemical and Biophysical Research Communications. 1999;260(1):273-279. DOI: 10.1006/bbrc.1999.0897
  48. 48. Ahuja N, Schwer B, Carobbio S, Waltregny D, North BJ, Castronovo V, et al. Regulation of insulin secretion by SIRT4, a mitochondrial ADP-ribosyltransferase. The Journal of Biological Chemistry. 2007;282(46):33583-33592. DOI: 10.1074/jbc.M705488200
  49. 49. Laurent G, German NJ, Saha AK, de Boer VC, Davies M, Koves TR, et al. SIRT4 coordinates the balance between lipid synthesis and catabolism by repressing malonyl COA decarboxylase. Molecular Cell. 2013;50(5):686-698. DOI: 10.1016/j.molcel.2013.05.012
  50. 50. Du J, Jiang H, Lin H. Investigating the ADP-ribosyltransferase activity of sirtuins with NAD analogues and 32p-NAD. Biochemistry. 2009;48(13):2878-2890. DOI: 10.1021/bi802093g
  51. 51. Butepage M, Eckei L, Verheugd P, Luscher B. Intracellular mono-ADP-ribosylation in signaling and disease. Cell. 2015;4(4):569-595. DOI: 10.3390/cells4040569
  52. 52. Haikarainen T, Lehtiö L. Proximal ADP-ribose hydrolysis in Trypanosomatids is catalyzed by a macrodomain. Scientific Reports. 2016;6:24213. DOI: 10.1038/srep24213
  53. 53. Jiang H, Khan S, Wang Y, Charron G, He B, Sebastian C, et al. SIRT6 regulates TNF-α secretion through hydrolysis of long-chain fatty acyl lysine. Nature. 2013;496(7443):110-113. DOI: 10.1038/nature12038
  54. 54. Park J, Chen Y, Tishkoff DX, Peng C, Tan M, Dai L, et al. SIRT5-mediated lysine desuccinylation impacts diverse metabolic pathways. Molecular Cell. 2013;50(6):919-930. DOI: 10.1016/j.molcel.2013.06.001
  55. 55. Nishida Y, Rardin MJ, Carrico C, He W, Sahu AK, Gut P, et al. SIRT5 regulates both cytosolic and mitochondrial protein malonylation with glycolysis as a major target. Molecular Cell. 2015;59(2):321-332. DOI: 10.1016/j.molcel.2015.05.022
  56. 56. Greiss S, Gartner A. Sirtuin/Sir2 phylogeny, evolutionary considerations and structural conservation. Molecules and Cells. 2009;28(5):407-415. DOI: 10.1007/s10059-009-0169-x
  57. 57. Chakrabarty SP, Balaram H. Reversible binding of zinc in Plasmodium falciparum Sir2: Structure and activity of the apoenzyme. Biochimica et Biophysica Acta. 2010;1804(9):1743-1750. DOI: 10.1016/j.bbapap.2010.06.010
  58. 58. Ritagliati C, Alonso VL, Manarin R, Cribb P, Serra EC. Overexpression of cytoplasmic TcSir2rp1 and mitochondrial TcSir2rp3 impacts on Trypanosoma cruzi growth and cell invasion. PLoS Neglected Tropical Diseases. 2015;9(4):e0003725. DOI: 10.1371/journal.pntd.0003725
  59. 59. Moretti NS, Augusto LD, Clemente TM, Antunes RP, Yoshida N, Torrecilhas AC, et al. Characterization of Trypanosoma cruzi sirtuins as possible drug targets for Chagas disease. Antimicrobial Agents and Chemotherapy. 2015;59(8):4669-4679. DOI: 10.1128/AAC.04694-14
  60. 60. Feldman JL, Dittenhafer-Reed KE, Denu JM. Sirtuin catalysis and regulation. The Journal of Biological Chemistry. 2012;287(51):42419-42427. DOI: 10.1074/jbc.R112.378877
  61. 61. Shore D, Squire M, Nasmyth KA. Characterization of two genes required for the position-effect control of yeast mating-type genes. The EMBO Journal. 1984;3(12):2817-2823
  62. 62. Ivy JM, Klar AJ, Hicks JB. Cloning and characterization of four Sir genes of Saccharomyces cerevisiae. Molecular and Cellular Biology. 1986;6(2):688-702
  63. 63. Aparicio OM, Billington BL, Gottschling DE. Modifiers of position effect are shared between telomeric and silent mating-type loci in S. cerevisiae. Cell. 1991;66(6):1279-1287
  64. 64. Bryk M, Banerjee M, Murphy M, Knudsen KE, Garfinkel DJ, Curcio MJ. Transcriptional silencing of Ty1 elements in the Rdn1 locus of yeast. Genes & Development. 1997;11(2):255-269
  65. 65. Fritze CE, Verschueren K, Strich R, Easton Esposito R. Direct evidence for Sir2 modulation of chromatin structure in yeast rDNA. The EMBO Journal. 1997;16(21):6495-6509. DOI: 10.1093/emboj/16.21.6495
  66. 66. Gottlieb S, Esposito RE. A new role for a yeast transcriptional silencer gene, Sir2, in regulation of recombination in ribosomal DNA. Cell. 1989;56(5):771-776
  67. 67. Shou W, Seol JH, Shevchenko A, Baskerville C, Moazed D, Chen ZW, et al. Exit from mitosis is triggered by TEM1-dependent release of the protein phosphatase cdc14 from nucleolar rent complex. Cell. 1999;97(2):233-244
  68. 68. Straight AF, Shou W, Dowd GJ, Turck CW, Deshaies RJ, Johnson AD, et al. Net1, a Sir2-associated nucleolar protein required for rDNA silencing and nucleolar integrity. Cell. 1999;97(2):245-256
  69. 69. Sinclair DA, Guarente L. Extrachromosomal rDNA circles-a cause of aging in yeast. Cell. 1997;91(7):1033-1042
  70. 70. Kaeberlein M, McVey M, Guarente L. The sir2/3/4 complex and sir2 alone promote longevity in saccharomyces cerevisiae by two different mechanisms. Genes & Development. 1999;13(19):2570-2580
  71. 71. Muth V, Nadaud S, Grummt I, Voit R. Acetylation of TAF(i)68, a subunit of tif-ib/sl1, activates RNA polymerase I transcription. The EMBO Journal. 2001;20(6):1353-1362. DOI: 10.1093/emboj/20.6.1353
  72. 72. Brooks CL, Gu W. Ubiquitination, phosphorylation and acetylation: The molecular basis for p53 regulation. Current Opinion in Cell Biology. 2003;15(2):164-171
  73. 73. Cheng HL, Mostoslavsky R, Saito S, Manis JP, Gu Y, Patel P, et al. 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(19):10794-10799. DOI: 10.1073/pnas.1934713100
  74. 74. Wang RH, Sengupta K, Li C, Kim HS, Cao L, Xiao C, et al. Impaired DNA damage response, genome instability, and tumorigenesis in SIRT1 mutant mice. Cancer Cell. 2008;14(4):312-323. DOI: 10.1016/j.ccr.2008.09.001
  75. 75. Canto C, Auwerx J. Caloric restriction, SIRT1 and longevity. Trends in Endocrinology and Metabolism. 2009;20(7):325-331. DOI: 10.1016/j.tem.2009.03.008
  76. 76. Hu Y, Liu J, Wang J, Liu Q. The controversial links among calorie restriction, SIRT1, and resveratrol. Free Radical Biology & Medicine. 2011;51(2):250-256. DOI: 10.1016/j.freeradbiomed.2011.04.034
  77. 77. Wang Y. Molecular links between caloric restriction and Sir2/SIRT1 activation. Diabetes and Metabolism Journal. 2014;38(5):321-329. DOI: 10.4093/dmj.2014.38.5.321
  78. 78. Kenyon CJ. The genetics of ageing. Nature. 2010;464(7288):504-512. DOI: 10.1038/nature08980
  79. 79. Fontana L, Partridge L, Longo VD. Extending healthy life span-from yeast to humans. Science. 2010;328(5976):321-326. DOI: 10.1126/science.1172539
  80. 80. Rodgers JT, Lerin C, Haas W, Gygi SP, Spiegelman BM, Puigserver P. Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature. 2005;434(7029):113-118. DOI: 10.1038/nature03354
  81. 81. Cohen HY, Miller C, Bitterman KJ, Wall NR, Hekking B, Kessler B, et al. Calorie restriction promotes mammalian cell survival by inducing the SIRT1 deacetylase. Science. 2004;305(5682):390-392. DOI: 10.1126/science.1099196
  82. 82. Civitarese AE, Carling S, Heilbronn LK, Hulver MH, Ukropcova B, Deutsch WA, et al. Calorie restriction increases muscle mitochondrial biogenesis in healthy humans. PLoS Medicine. 2007;4(3):485-494. DOI: 10.1371/journal.pmed.0040076
  83. 83. Boily G, Seifert EL, Bevilacqua L, He XH, Sabourin G, Estey C, et al. SIRT1 regulates energy metabolism and response to caloric restriction in mice. PLoS One. 2008;3(3):e1759. DOI: 10.1371/journal.pone.0001759
  84. 84. North BJ, Verdin E. Mitotic regulation of SIRT2 by cyclin-dependent kinase 1-dependent phosphorylation. The Journal of Biological Chemistry. 2007;282(27):19546-19555. DOI: 10.1074/jbc.M702990200
  85. 85. Vaquero A, Scher MB, Lee DH, Sutton A, Cheng HL, Alt FW, et al. SIRT2 is a histone deacetylase with preference for histone h4 lys 16 during mitosis. Genes & Development. 2006;20(10):1256-1261. DOI: 10.1101/gad.1412706
  86. 86. Dryden SC, Nahhas FA, Nowak JE, Goustin AS, Tainsky MA. Role for human SIRT2 NAD-dependent deacetylase activity in control of mitotic exit in the cell cycle. Molecular and Cellular Biology. 2003;23(9):3173-3185
  87. 87. Suzuki K, Koike T. Mammalian Sir2-related protein (SIRT) 2-mediated modulation of resistance to axonal degeneration in slow wallerian degeneration mice: A crucial role of tubulin deacetylation. Neuroscience. 2007;147(3):599-612. DOI: 10.1016/j.neuroscience.2007.04.059
  88. 88. Luthi-Carter R, Taylor DM, Pallos J, Lambert E, Amore A, Parker A, et al. SIRT2 inhibition achieves neuroprotection by decreasing sterol biosynthesis. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(17):7927-7932. DOI: 10.1073/pnas.1002924107
  89. 89. Outeiro TF, Kontopoulos E, Altmann SM, Kufareva I, Strathearn KE, Amore AM, et al. Sirtuin 2 inhibitors rescue alpha-synuclein-mediated toxicity in models of parkinson's disease. Science. 2007;317(5837):516-519. DOI: 10.1126/science.1143780
  90. 90. Kim HS, Vassilopoulos A, Wang RH, Lahusen T, Xiao Z, Xu X, et al. SIRT2 maintains genome integrity and suppresses tumorigenesis through regulating apc/c activity. Cancer Cell. 2011;20(4):487-499. DOI: 10.1016/j.ccr.2011.09.004
  91. 91. Hiratsuka M, Inoue T, Toda T, Kimura N, Shirayoshi Y, Kamitani H, et al. Proteomics-based identification of differentially expressed genes in human gliomas: Down-regulation of SIRT2 gene. Biochemical and Biophysical Research Communications. 2003;309(3):558-566
  92. 92. Ahn BH, Kim HS, Song S, Lee IH, Liu J, Vassilopoulos A, et al. A role for the mitochondrial deacetylase SIRT3 in regulating energy homeostasis. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(38):14447-14452. DOI: 10.1073/pnas.0803790105
  93. 93. Lombard DB, Alt FW, Cheng HL, Bunkenborg J, Streeper RS, Mostoslavsky R, et al. Mammalian Sir2 homolog SIRT3 regulates global mitochondrial lysine acetylation. Molecular and Cellular Biology. 2007;27(24):8807-8814. DOI: 10.1128/MCB.01636-07
  94. 94. Hirschey MD, Shimazu T, Goetzman E, Jing E, Schwer B, Lombard DB, et al. SIRT3 regulates mitochondrial fatty-acid oxidation by reversible enzyme deacetylation. Nature. 2010;464(7285):121-125. DOI: 10.1038/nature08778
  95. 95. Shimazu T, Hirschey MD, Hua L, Dittenhafer-Reed KE, Schwer B, Lombard DB, et al. SIRT3 deacetylates mitochondrial 3-hydroxy-3-methylglutaryl COA synthase 2 and regulates ketone body production. Cell Metabolism. 2010;12(6):654-661. DOI: 10.1016/j.cmet.2010.11.003
  96. 96. Onyango P, Celic I, McCaffery JM, Boeke JD, Feinberg AP. SIRT3, a human sir2 homologue, is an NAD-dependent deacetylase localized to mitochondria. Proceedings of the National Academy of Sciences of the United States of America. 2002;99(21):13653-13658. DOI: 10.1073/pnas.222538099
  97. 97. Schwer B, North BJ, Frye RA, Ott M, Verdin E. The human silent information regulator (Sir)2 homologue hSIRT3 is a mitochondrial nicotinamide adenine dinucleotide-dependent deacetylase. The Journal of Cell Biology. 2002;158(4):647-657. DOI: 10.1083/jcb.200205057
  98. 98. Shi T, Wang F, Stieren E, Tong Q. SIRT3, a mitochondrial sirtuin deacetylase, regulates mitochondrial function and thermogenesis in brown adipocytes. The Journal of Biological Chemistry. 2005;280(14):13560-13567. DOI: 10.1074/jbc.M414670200
  99. 99. He W, Newman JC, Wang MZ, Ho L, Verdin E. Mitochondrial sirtuins: Regulators of protein acylation and metabolism. Trends in Endocrinology and Metabolism. 2012;23(9):467-476. DOI: 10.1016/j.tem.2012.07.004
  100. 100. Kincaid B, Bossy-Wetzel E. Forever young: SIRT3 a shield against mitochondrial meltdown, aging, and neurodegeneration. Frontiers in Aging Neuroscience. 2013;5:48-61. DOI: 10.3389/fnagi.2013.00048
  101. 101. Haigis MC, Mostoslavsky R, Haigis KM, Fahie K, Christodoulou DC, Murphy AJ, et al. SIRT4 inhibits glutamate dehydrogenase and opposes the effects of calorie restriction in pancreatic beta cells. Cell. 2006;126(5):941-954. DOI: 10.1016/j.cell.2006.06.057
  102. 102. Jeong SM, Xiao C, Finley LW, Lahusen T, Souza AL, Pierce K, et al. SIRT4 has tumor-suppressive activity and regulates the cellular metabolic response to DNA damage by inhibiting mitochondrial glutamine metabolism. Cancer Cell. 2013;23(4):450-463. DOI: 10.1016/j.ccr.2013.02.024
  103. 103. Csibi A, Fendt SM, Li C, Poulogiannis G, Choo AY, Chapski DJ, et al. The mTORc1 pathway stimulates glutamine metabolism and cell proliferation by repressing SIRT4. Cell. 2013;153(4):840-854. DOI: 10.1016/j.cell.2013.04.023
  104. 104. Du J, Zhou Y, Su X, Yu JJ, Khan S, Jiang H, et al. SIRT5 is a NAD-dependent protein lysine demalonylase and desuccinylase. Science. 2011;334(6057):806-809. DOI: 10.1126/science.1207861
  105. 105. Tan M, Peng C, Anderson KA, Chhoy P, Xie Z, Dai L, et al. Lysine glutarylation is a protein posttranslational modification regulated by SIRT5. Cell Metabolism. 2014;19(4):605-617. DOI: 10.1016/j.cmet.2014.03.014
  106. 106. Nakamura Y, Ogura M, Ogura K, Tanaka D, Inagaki N. SIRT5 deacetylates and activates urate oxidase in liver mitochondria of mice. FEBS Letters. 2012;586(23):4076-4081. DOI: 10.1016/j.febslet.2012.10.009
  107. 107. Yu J, Sadhukhan S, Noriega LG, Moullan N, He B, Weiss RS, et al. Metabolic characterization of a SIRT5 deficient mouse model. Scientific Reports. 2013;3:2806-2813. DOI: 10.1038/srep02806
  108. 108. Kanfi Y, Peshti V, Gil R, Naiman S, Nahum L, Levin E, et al. SIRT6 protects against pathological damage caused by diet-induced obesity. Aging Cell. 2010;9(2):162-173. DOI: 10.1111/j.1474-9726.2009.00544.x
  109. 109. Mostoslavsky R, Chua KF, Lombard DB, Pang WW, Fischer MR, Gellon L, et al. Genomic instability and aging-like phenotype in the absence of mammalian SIRT6. Cell. 2006;124(2):315-329. DOI: 10.1016/j.cell.2005.11.044
  110. 110. Liszt G, Ford E, Kurtev M, Guarente L. Mouse Sir2 homolog SIRT6 is a nuclear ADP-ribosyltransferase. The Journal of Biological Chemistry. 2005;280(22):21313-21320. DOI: 10.1074/jbc.M413296200
  111. 111. Michishita E, McCord RA, Berber E, Kioi M, Padilla-Nash H, Damian M, et al. SIRT6 is a histone H3 lysine 9 deacetylase that modulates telomeric chromatin. Nature. 2008;452(7186):492-496. DOI: 10.1038/nature06736
  112. 112. Michishita E, McCord RA, Boxer LD, Barber MF, Hong T, Gozani O, et al. Cell cycle-dependent deacetylation of telomeric histone h3 lysine k56 by human SIRT6. Cell Cycle. 2009;8(16):2664-2666. DOI: 10.4161/cc.8.16.9367
  113. 113. Yang B, Zwaans BM, Eckersdorff M, Lombard DB. The sirtuin SIRT6 deacetylates H3 K56Ac in vivo to promote genomic stability. Cell Cycle. 2009;8(16):2662-2663. DOI: 10.4161/cc.8.16.9329
  114. 114. Mao Z, Hine C, Tian X, Van Meter M, Au M, Vaidya A, et al. SIRT6 promotes DNA repair under stress by activating PARP1. Science. 2011;332(6036):1443-1446. DOI: 10.1126/science.1202723
  115. 115. Mao Z, Tian X, Van Meter M, Ke Z, Gorbunova V, Seluanov A. Sirtuin 6 (SIRT6) rescues the decline of homologous recombination repair during replicative senescence. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(29):11800-11805. DOI: 10.1073/pnas.1200583109
  116. 116. Zhong L, D'Urso A, Toiber D, Sebastian C, Henry RE, Vadysirisack DD, et al. The histone deacetylase SIRT6 regulates glucose homeostasis via HIF1alpha. Cell. 2010;140(2):280-293. DOI: 10.1016/j.cell.2009.12.041
  117. 117. Dominy JE Jr, Lee Y, Jedrychowski MP, Chim H, Jurczak MJ, Camporez JP, et al. The deacetylase SIRT6 activates the acetyltransferase GCN5 and suppresses hepatic gluconeogenesis. Molecular Cell. 2012;48(6):900-913. DOI: 10.1016/j.molcel.2012.09.030
  118. 118. Puigserver P, Rhee J, Donovan J, Walkey CJ, Yoon JC, Oriente F, et al. Insulin-regulated hepatic gluconeogenesis through FOXO1-PGC-1α interaction. Nature. 2003;423(6939):550-555. DOI: 10.1038/nature01667
  119. 119. Kim HS, Xiao C, Wang RH, Lahusen T, Xu X, Vassilopoulos A, et al. Hepatic-specific disruption of SIRT6 in mice results in fatty liver formation due to enhanced glycolysis and triglyceride synthesis. Cell Metabolism. 2010;12(3):224-236. DOI: 10.1016/j.cmet.2010.06.009
  120. 120. Etchegaray JP, Zhong L, Mostoslavsky R. The histone deacetylase SIRT6: At the crossroads between epigenetics, metabolism and disease. Current Topics in Medicinal Chemistry. 2013;13(23):2991-3000
  121. 121. Michishita E, Park JY, Burneskis JM, Barrett JC, Horikawa I. Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Molecular Biology of the Cell. 2005;16(10):4623-4635. DOI: 10.1091/mbc.E05-01-0033
  122. 122. Ford E, Voit R, Liszt G, Magin C, Grummt I, Guarente L. Mammalian Sir2 homolog SIRT7 is an activator of RNAa polymerase I transcription. Genes & Development. 2006;20(9):1075-1080. DOI: 10.1101/gad.1399706
  123. 123. Tsai YC, Greco TM, Cristea IM. Sirtuin 7 plays a role in ribosome biogenesis and protein synthesis. Molecular & Cellular Proteomics. 2014;13(1):73-83. DOI: 10.1074/mcp.M113.031377
  124. 124. Karim MF, Yoshizawa T, Sato Y, Sawa T, Tomizawa K, Akaike T, et al. Inhibition of H3K18 deacetylation of SIRT7 by MYB-binding protein 1a (MYBBP1a). Biochemical and Biophysical Research Communications. 2013;441(1):157-163. DOI: 10.1016/j.bbrc.2013.10.020
  125. 125. Tsai YC, Greco TM, Boonmee A, Miteva Y, Cristea IM. Functional proteomics establishes the interaction of SIRT7 with chromatin remodeling complexes and expands its role in regulation of RNA polymerase i transcription. Molecular & Cellular Proteomics. 2012;11(5):60-76. DOI: 10.1074/mcp.A111.015156
  126. 126. Kim JK, Noh JH, Jung KH, Eun JW, Bae HJ, Kim MG, et al. Sirtuin7 oncogenic potential in human hepatocellular carcinoma and its regulation by the tumor suppressors mir-125a-5p and mir-125b. Hepatology. 2013;57(3):1055-1067. DOI: 10.1002/hep.26101
  127. 127. Vakhrusheva O, Braeuer D, Liu Z, Braun T, Bober E. SIRT7-dependent inhibition of cell growth and proliferation might be instrumental to mediate tissue integrity during aging. Journal of Physiology and Pharmacology. 2008;59(Suppl 9):201-212
  128. 128. Kiran S, Oddi V, Ramakrishna G. Sirtuin 7 promotes cellular survival following genomic stress by attenuation of DNA damage, sapk activation and p53 response. Experimental Cell Research. 2015;331(1):123-141. DOI: 10.1016/j.yexcr.2014.11.001
  129. 129. Chen S, Seiler J, Santiago-Reichelt M, Felbel K, Grummt I, Voit R. Repression of RNA polymerase I upon stress is caused by inhibition of RNA-dependent deacetylation of PAF53 by SIRT7. Molecular Cell. 2013;52(3):303-313. DOI: 10.1016/j.molcel.2013.10.010
  130. 130. Hubbi ME, Hu H, Kshitiz, Gilkes DM, Semenza GL. Sirtuin-7 inhibits the activity of hypoxia-inducible factors. The Journal of Biological Chemistry. 2013;288(29):20768-20775. DOI: 10.1074/jbc.M113.476903
  131. 131. Ashraf N, Zino S, Macintyre A, Kingsmore D, Payne AP, George WD, et al. Altered sirtuin expression is associated with node-positive breast cancer. British Journal of Cancer. 2006;95(8):1056-1061. DOI: 10.1038/sj.bjc.6603384
  132. 132. Han Y, Liu Y, Zhang H, Wang T, Diao R, Jiang Z, et al. Hsa-mir-125b suppresses bladder cancer development by down-regulating oncogene SIRT7 and oncogenic long noncoding RNA malat1. FEBS Letter. 2013; S0014-5793(13)00780-1. DOI: 10.1016/j.febslet.2013.10.023
  133. 133. Freitas-Junior LH, Hernandez-Rivas R, Ralph SA, Montiel-Condado D, Ruvalcaba-Salazar OK, Rojas-Meza AP, et al. Telomeric heterochromatin propagation and histone acetylation control mutually exclusive expression of antigenic variation genes in malaria parasites. Cell. 2005;121(1):25-36. DOI: 10.1016/j.cell.2005.01.037
  134. 134. Issar N, Roux E, Mattei D, Scherf A. Identification of a novel post-translational modification in Plasmodium falciparum: Protein sumoylation in different cellular compartments. Cellular Microbiology. 2008;10(10):1999-2011. DOI: 10.1111/j.1462-5822.2008.01183.x
  135. 135. Duraisingh MT, Voss TS, Marty AJ, Duffy MF, Good RT, Thompson JK, et al. Heterochromatin silencing and locus repositioning linked to regulation of virulence genes in Plasmodium falciparum. Cell. 2005;121(1):13-24. DOI: 10.1016/j.cell.2005.01.036
  136. 136. Hviid L, Jensen AT. Pfemp1—A parasite protein family of key importance in Plasmodium falciparum malaria immunity and pathogenesis. Advances in Parasitology. 2015;88:51-84. DOI: 10.1016/bs.apar.2015.02.004
  137. 137. Scherf A, Hernandez-Rivas R, Buffet P, Bottius E, Benatar C, Pouvelle B, et al. Antigenic variation in malaria: In situ switching, relaxed and mutually exclusive transcription of var genes during intra-erythrocytic development in Plasmodium falciparum. The EMBO Journal. 1998;17(18):5418-5426. DOI: 10.1093/emboj/17.18.5418
  138. 138. Chookajorn T, Ponsuwanna P, Cui L. Mutually exclusive var gene expression in the malaria parasite: Multiple layers of regulation. Trends in Parasitology. 2008;24(10):455-461. DOI: 10.1016/j.pt.2008.07.005
  139. 139. Deitsch KW, del Pinal A, Wellems TE. Intra-cluster recombination and var transcription switches in the antigenic variation of Plasmodium falciparum. Molecular and Biochemical Parasitology. 1999;101(1–2):107-116
  140. 140. Tonkin CJ, Carret CK, Duraisingh MT, Voss TS, Ralph SA, Hommel M, et al. Sir2 paralogues cooperate to regulate virulence genes and antigenic variation in Plasmodium falciparum. PLoS Biology. 2009;7(4):e84. DOI: 10.1371/journal.pbio.1000084
  141. 141. Yahiaoui B, Taibi A, Ouaissi A. A Leishmania major protein with extensive homology to silent information regulator 2 of saccharomyces cerevisiae. Gene. 1996;169(1):115-118
  142. 142. Zemzoumi K, Sereno D, Francois C, Guilvard E, Lemesre JL, Ouaissi A. Leishmania major: Cell type dependent distribution of a 43 KDa antigen related to silent information regulatory-2 protein family. Biology of the Cell. 1998;90(3):239-245. DOI: 10.1016/S0248-4900(98)80020-8
  143. 143. Vergnes B, Sereno D, Madjidian-Sereno N, Lemesre JL, Ouaissi A. Cytoplasmic Sir2 homologue overexpression promotes survival of Leishmania parasites by preventing programmed cell death. Gene. 2002;296(1–2):139-150
  144. 144. Sereno D, Vanhille L, Vergnes B, Monte-Allegre A, Ouaissi A. Experimental study of the function of the excreted/secreted Leishmania LmSir2 protein by heterologous expression in eukaryotic cell line. Kinetoplastid Biology and Disease. 2005;4(1):1-9. DOI: 10.1186/1475-9292-4-1
  145. 145. Vergnes B, Sereno D, Tavares J, Cordeiro-da-Silva A, Vanhille L, Madjidian-Sereno N, et al. Targeted disruption of cytosolic Sir2 deacetylase discloses its essential role in Leishmania survival and proliferation. Gene. 2005;363:85-96. DOI: 10.1016/j.gene.2005.06.047
  146. 146. Tavares J, Ouaissi A, Santarem N, Sereno D, Vergnes B, Sampaio P, et al. The Leishmania infantum cytosolic Sir2-related protein 1 (lisir2rp1) is an NAD+-dependent deacetylase and ADP-ribosyltransferase. The Biochemical Journal. 2008;415(3):377-386. DOI: 10.1042/BJ20080666
  147. 147. Adriano MA, Vergnes B, Poncet J, Mathieu-Daude F, da Silva AC, Ouaissi A, et al. Proof of interaction between Leishmania Sir2rp1 deacetylase and chaperone hsp83. Parasitology Research 2007;100(4):811-818. DOI: 10.1007/s00436-006-0352-3
  148. 148. Matsuyama A, Shimazu T, Sumida Y, Saito A, Yoshimatsu Y, Seigneurin-Berny D, et al. In vivo destabilization of dynamic microtubules by HDAC6-mediated deacetylation. The EMBO Journal. 2002;21(24):6820-6831. DOI: 10.1093/emboj/cdf682
  149. 149. Fessel MR, Lira CB, Giorgio S, Ramos CH, Cano MI. Sir2-related protein 1 from Leishmania amazonensis is a glycosylated NAD+-dependent deacetylase. Parasitology. 2011;138(10):1245-1258. DOI: 10.1017/S0031182011001077
  150. 150. Purkait B, Singh R, Wasnik K, Das S, Kumar A, Paine M, et al. Up-regulation of silent information regulator 2 (Sir2) is associated with amphotericin B resistance in clinical isolates of Leishmania donovani. The Journal of Antimicrobial Chemotherapy. 2015;70(5):1343-1356. DOI: 10.1093/jac/dku534
  151. 151. Mishra J, Singh S. Miltefosine resistance in Leishmania donovani involves suppression of oxidative stress-induced programmed cell death. Experimental Parasitology. 2013;135(2):397-406. DOI: 10.1016/j.exppara.2013.08.004
  152. 152. Silvestre R, Cordeiro-Da-Silva A, Santarem N, Vergnes B, Sereno D, Ouaissi A. Sir2-deficient Leishmania infantum induces a defined IFN-gamma/IL-10 pattern that correlates with protection. Journal of Immunology. 2007;179(5):3161-3170
  153. 153. Baharia RK, Tandon R, Sharma T, Suthar MK, Das S, Siddiqi MI, et al. Recombinant NAD-dependent sir-2 protein of Leishmania donovani: Immunobiochemical characterization as a potential vaccine against visceral Leishmaniasis. PLoS Neglected Tropical Diseases. 2015;9(3):e0003557. DOI: 10.1371/journal.pntd.0003557
  154. 154. Vergnes B, Gazanion E, Grentzinger T. Functional divergence of Sir2 orthologs between Trypanosomatid parasites. Molecular and Biochemical Parasitology. 2016;207(2):96-101. DOI: 10.1016/j.molbiopara.2016.06.004
  155. 155. Garcia-Salcedo JA, Gijon P, Nolan DP, Tebabi P, Pays E. A chromosomal Sir2 homologue with both histone NAD-dependent ADP-ribosyltransferase and deacetylase activities is involved in DNA repair in Trypanosoma brucei. The EMBO Journal. 2003;22(21):5851-5862. DOI: 10.1093/emboj/cdg553
  156. 156. Brachmann CB, Sherman JM, Devine SE, Cameron EE, Pillus L, Boeke JD. The Sir2 gene family, conserved from bacteria to humans, functions in silencing, cell cycle progression, and chromosome stability. Genes & Development. 1995;9(23):2888-2902
  157. 157. Landry J, Sutton A, Tafrov ST, Heller RC, Stebbins J, Pillus L, et al. The silencing protein Sir2 and its homologs are NAD-dependent protein deacetylases. Proceedings of the National Academy of Sciences of the United States of America. 2000;97(11):5807-5811. DOI: 10.1073/pnas.110148297
  158. 158. Perrod S, Cockell MM, Laroche T, Renauld H, Ducrest AL, Bonnard C, et al. A cytosolic NAD-dependent deacetylase, HST2p, can modulate nucleolar and telomeric silencing in yeast. The EMBO Journal. 2001;20(1–2):197-209. DOI: 10.1093/emboj/20.1.197
  159. 159. Tanner KG, Landry J, Sternglanz R, Denu JM. Silent information regulator 2 family of NAD-dependent histone/protein deacetylases generates a unique product, 1-O-acetyl-ADP-ribose. Proceedings of the National Academy of Sciences of the United States of America. 2000;97(26):14178-14182. DOI: 10.1073/pnas.250422697
  160. 160. Alsford S, Kawahara T, Isamah C, Horn D. A sirtuin in the african trypanosome is involved in both DNA repair and telomeric gene silencing but is not required for antigenic variation. Molecular Microbiology. 2007;63(3):724-736. DOI: 10.1111/j.1365-2958.2006.05553.x
  161. 161. Rusche LN, Kirchmaier AL, Rine J. The establishment, inheritance, and function of silenced chromatin in Saccharomyces cerevisiae. Annual Review of Biochemistry. 2003;72:481-516. DOI: 10.1146/annurev.biochem.72.121801.161547
  162. 162. Figueiredo L, Scherf A. Plasmodium telomeres and telomerase: The usual actors in an unusual scenario. Chromosome Research. 2005;13(5):517-524. DOI: 10.1007/s10577-005-0996-3
  163. 163. Kowieski TM, Lee S, Denu JM. Acetylation-dependent ADP-ribosylation by Trypanosoma brucei Sir2. The Journal of Biological Chemistry. 2008;283(9):5317-5326. DOI: 10.1074/jbc.M707613200
  164. 164. Fahie K, Hu P, Swatkoski S, Cotter RJ, Zhang Y, Wolberger C. Side chain specificity of ADP-ribosylation by a sirtuin. The FEBS Journal. 2009;276(23):7159-7176. DOI: 10.1111/j.1742-4658.2009.07427.x
  165. 165. Hailu GS, Robaa D, Forgione M, Sippl W, Rotili D, Mai A. Lysine deacetylase inhibitors in parasites: Past, present, and future perspectives. Journal of Medicinal Chemistry. 2017;60(12):4780-4804. DOI: 10.1021/acs.jmedchem.6b01595
  166. 166. Morselli E, Maiuri MC, Markaki M, Megalou E, Pasparaki A, Palikaras K, et al. Caloric restriction and resveratrol promote longevity through the sirtuin-1-dependent induction of autophagy. Cell Death & Disease. 2010;1:e10. DOI: 10.1038/cddis.2009.8
  167. 167. Guarente L, Picard F. Calorie restriction-the Sir2 connection. Cell. 2005;120(4):473-482. DOI: 10.1016/j.cell.2005.01.029
  168. 168. Lara E, Mai A, Calvanese V, Altucci L, Lopez-Nieva P, Martinez-Chantar ML, et al. Salermide, a sirtuin inhibitor with a strong cancer-specific proapoptotic effect. Oncogene. 2009;28(6):781-791. DOI: 10.1038/onc.2008.436
  169. 169. Zheng W. Sirtuins as emerging anti-parasitic targets. European Journal of Medicinal Chemistry. 2013;59:132-140. DOI: 10.1016/j.ejmech.2012.11.014
  170. 170. Religa AA, Waters AP. Sirtuins of parasitic protozoa: In search of function(s). Molecular and Biochemical Parasitology. 2012;185(2):71-88. DOI: 10.1016/j.molbiopara.2012.08.003
  171. 171. Rogina B, Helfand SL, Frankel S. Longevity regulation by Drosophila RPD3 deacetylase and caloric restriction. Science. 2002;298(5599):1745. DOI: 10.1126/science.1078986
  172. 172. Tissenbaum HA, Guarente L. Increased dosage of a Sir-2 gene extends lifespan in Caenorhabditis elegans. Nature. 2001;410(6825):227-230. DOI: 10.1038/35065638
  173. 173. Eddy SR. Where did the BLOSUM62 alignment score matrix come from? Nature Biotechnology. 2004;22(8):1035-1036. DOI: 10.1038/nbt0804-1035
  174. 174. Hajduk PJ, Huth JR, Tse C. Predicting protein druggability. Drug Discovery Today. 2005;10(23–24):1675-1682. DOI: 10.1016/S1359-6446(05)03624-X
  175. 175. Tavares J, Ouaissi A, Kong Thoo Lin P, Loureiro I, Kaur S, Roy N, et al. Bisnaphthalimidopropyl derivatives as inhibitors of Leishmania Sir2 related protein 1. ChemMedChem. 2010;5(1):140-147. DOI: 10.1002/cmdc.200900367
  176. 176. Schlicker C, Boanca G, Lakshminarasimhan M, Steegborn C. Structure-based development of novel sirtuin inhibitors. Aging (Albany NY). 2011;3(9):852-872
  177. 177. Rumpf T, Schiedel M, Karaman B, Roessler C, North BJ, Lehotzky A, et al. Selective SIRT2 inhibition by ligand-induced rearrangement of the active site. Nature Communications. 2015;6:6263-6276. DOI: 10.1038/ncomms7263
  178. 178. Avalos JL, Bever KM, Wolberger C. Mechanism of sirtuin inhibition by nicotinamide: Altering the NAD(+) cosubstrate specificity of a Sir2 enzyme. Molecular Cell. 2005;17(6):855-868. DOI: 10.1016/j.molcel.2005.02.022
  179. 179. Soares MB, Silva CV, Bastos TM, Guimaraes ET, Figueira CP, Smirlis D, et al. Anti-Trypanosoma cruzi activity of nicotinamide. Acta Tropica. 2012;122(2):224-229. DOI: 10.1016/j.actatropica.2012.01.001
  180. 180. Prusty D, Mehra P, Srivastava S, Shivange AV, Gupta A, Roy N, et al. Nicotinamide inhibits Plasmodium falciparum Sir2 activity in vitro and parasite growth. FEMS Microbiology Letters. 2008;282(2):266-272. DOI: 10.1111/j.1574-6968.2008.01135.x
  181. 181. Sereno D, Alegre AM, Silvestre R, Vergnes B, Ouaissi A. In vitro antiLeishmanial activity of nicotinamide. Antimicrobial Agents and Chemotherapy. 2005;49(2):808-812. DOI: 10.1128/AAC.49.2.808-812.2005
  182. 182. Unciti-Broceta JD, Maceira J, Morales S, Garcia-Perez A, Munoz-Torres ME, Garcia-Salcedo JA. Nicotinamide inhibits the lysosomal cathepsin b-like protease and kills African trypanosomes. The Journal of Biological Chemistry. 2013;288(15):10548-10557. DOI: 10.1074/jbc.M112.449207
  183. 183. Corda D, Di Girolamo M. Functional aspects of protein mono-ADP-ribosylation. The EMBO Journal. 2003;22(9):1953-1958. DOI: 10.1093/emboj/cdg209
  184. 184. Ludden PW. Reversible ADP-ribosylation as a mechanism of enzyme regulation in procaryotes. Molecular and Cellular Biochemistry. 1994;138(1–2):123-129
  185. 185. Hemphill A, Lawson D, Seebeck T. The cytoskeletal architecture of Trypanosoma brucei. The Journal of Parasitology. 1991;77(4):603-612
  186. 186. Inoue T, Hiratsuka M, Osaki M, Yamada H, Kishimoto I, Yamaguchi S, et al. SIRT2, a tubulin deacetylase, acts to block the entry to chromosome condensation in response to mitotic stress. Oncogene. 2007;26(7):945-957. DOI: 10.1038/sj.onc.1209857
  187. 187. Pandithage R, Lilischkis R, Harting K, Wolf A, Jedamzik B, Luscher-Firzlaff J, et al. The regulation of SIRT2 function by cyclin-dependent kinases affects cell motility. The Journal of Cell Biology. 2008;180(5):915-929. DOI: 10.1083/jcb.200707126
  188. 188. Tanno M, Sakamoto J, Miura T, Shimamoto K, Horio Y. Nucleocytoplasmic shuttling of the NAD+-dependent histone deacetylase SIRT1. The Journal of Biological Chemistry. 2007;282(9):6823-6832. DOI: 10.1074/jbc.M609554200
  189. 189. Cho Y, Sloutsky R, Naegle KM, Cavalli V. Injury-induced HDAC5 nuclear export is essential for axon regeneration. Cell. 2013;155(4):894-908. DOI: 10.1016/j.cell.2013.10.004
  190. 190. North BJ, Verdin E. Interphase nucleo-cytoplasmic shuttling and localization of SIRT2 during mitosis. PLoS One. 2007;2(8):e784. DOI: 10.1371/journal.pone.0000784
  191. 191. Prieto G, Fullaondo A, Rodriguez JA. Prediction of nuclear export signals using weighted regular expressions (Wregex). Bioinformatics. 2014;30(9):1220-1227. DOI: 10.1093/bioinformatics/btu016
  192. 192. Kosugi S, Hasebe M, Tomita M, Yanagawa H. Systematic identification of cell cycle-dependent yeast nucleocytoplasmic shuttling proteins by prediction of composite motifs. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(25):10171-10176. DOI: 10.1073/pnas.0900604106
  193. 193. Kamal A, Bolla NR, Srikanth PS, Srivastava AK. Naphthalimide derivatives with therapeutic characteristics: A patent review. Expert Opinion on Therapeutic Patents. 2013;23(3):299-317. DOI: 10.1517/13543776.2013.746313
  194. 194. Filosa R, Peduto A, Micco SD, Caprariis P, Festa M, Petrella A, et al. Molecular modelling studies, synthesis and biological activity of a series of novel bisnaphthalimides and their development as new DNA topoisomerase II inhibitors. Bioorganic & Medicinal Chemistry. 2009;17(1):13-24. DOI: 10.1016/j.bmc.2008.11.024
  195. 195. Seliga R, Pilatova M, Sarissky M, Viglasky V, Walko M, Mojzis J. Novel naphthalimide polyamine derivatives as potential antitumor agents. Molecular Biology Reports. 2013;40(6):4129-4137. DOI: 10.1007/s11033-013-2523-5
  196. 196. Wang X, Chen Z, Tong L, Tan S, Zhou W, Peng T, et al. Naphthalimides exhibit in vitro antiproliferative and antiangiogenic activities by inhibiting both topoisomerase II (topo II) and receptor tyrosine kinases (rtks). European Journal of Medicinal Chemistry. 2013;65:477-486. DOI: 10.1016/j.ejmech.2013.05.002
  197. 197. Costanza ME, Berry D, Henderson IC, Ratain MJ, Wu K, Shapiro C, et al. Amonafide: An active agent in the treatment of previously untreated advanced breast cancer--a cancer and leukemia group b study (calgb 8642). Clinical Cancer Research. 1995;1(7):699-704
  198. 198. Thompson J, Pratt CB, Stewart CF, Avery L, Bowman L, Zamboni WC, et al. Phase I study of DMP 840 in pediatric patients with refractory solid tumors. Investigational New Drugs. 1998;16(1):45-49
  199. 199. Bousquet PF, Brana MF, Conlon D, Fitzgerald KM, Perron D, Cocchiaro C, et al. Preclinical evaluation of LU 79553: A novel bis-naphthalimide with potent antitumor activity. Cancer Research. 1995;55(5):1176-1180
  200. 200. Dance AM, Ralton L, Fuller Z, Milne L, Duthie S, Bestwick CS, et al. Synthesis and biological activities of bisnaphthalimido polyamines derivatives: Cytotoxicity, DNA binding, DNA damage and drug localization in breast cancer MCF 7 cells. Biochemical Pharmacology. 2005;69(1):19-27. DOI: 10.1016/j.bcp.2004.09.020
  201. 201. Oliveira J, Ralton L, Tavares J, Codeiro-da-Silva A, Bestwick CS, McPherson A, et al. The synthesis and the in vitro cytotoxicity studies of bisnaphthalimidopropyl polyamine derivatives against colon cancer cells and parasite Leishmania infantum. Bioorganic & Medicinal Chemistry. 2007;15(1):541-545. DOI: 10.1016/j.bmc.2006.09.031
  202. 202. Graca NA, Gaspar L, Costa DM, Loureiro I, Thoo-Lin PK, Ramos I, et al. Activity of bisnaphthalimidopropyl derivatives against Trypanosoma brucei. Antimicrobial Agents and Chemotherapy. 2016;60(4):2532-2536. DOI: 10.1128/aac.02490-15
  203. 203. Ralton L, Bestwick CS, Thoo Lin PK. Polyamine analogues and derivatives as potential anticancer agents. Current Bioactive Compounds. 2007;3(3):179-191
  204. 204. Braña MF, Castellano JM, Morán M, Pérez de Vega MJ, Qian XD, Romerdahl CA, et al. Bis-naphthalimides. 2. Synthesis and biological activity of 5,6-acenaphthalimidoalkyl-1,8-naphthalimidoalkyl amines. European Journal of Medicinal Chemistry. 1995;30(3):235-239. DOI: 10.1016/0223-5234(96)88230-4
  205. 205. Roy Chowdhury A, Bakshi R, Wang J, Yildirir G, Liu B, Pappas-Brown V, et al. The killing of african trypanosomes by ethidium bromide. PLoS Pathogens. 2010;6(12):e1001226. DOI: 10.1371/journal.ppat.1001226
  206. 206. Terstappen GC, Schlupen C, Raggiaschi R, Gaviraghi G. Target deconvolution strategies in drug discovery. Nature Reviews. Drug Discovery. 2007;6(11):891-903. DOI: 10.1038/nrd2410
  207. 207. Hirota T, Lee JW, St John PC, Sawa M, Iwaisako K, Noguchi T, et al. Identification of small molecule activators of cryptochrome. Science. 2012;337(6098):1094-1097. DOI: 10.1126/science.1223710
  208. 208. Schenone M, Dancik V, Wagner BK, Clemons PA. Target identification and mechanism of action in chemical biology and drug discovery. Nature Chemical Biology. 2013;9(4):232-240. DOI: 10.1038/nchembio.1199
  209. 209. Wyllie S, Oza SL, Patterson S, Spinks D, Thompson S, Fairlamb AH. Dissecting the essentiality of the bifunctional trypanothione synthetase-amidase in Trypanosoma brucei using chemical and genetic methods. Molecular Microbiology. 2009;74(3):529-540. DOI: 10.1111/j.1365-2958.2009.06761.x
  210. 210. Roberts AJ, Torrie LS, Wyllie S, Fairlamb AH. Biochemical and genetic characterization of Trypanosoma cruzi N-myristoyltransferase. The Biochemical Journal. 2014;459(2):323-332. DOI: 10.1042/BJ20131033
  211. 211. Jones DC, Foth BJ, Urbaniak MD, Patterson S, Ong HB, Berriman M, et al. Genomic and proteomic studies on the mode of action of oxaboroles against the african trypanosome. PLoS Neglected Tropical Diseases. 2015;9(12):e0004299. DOI: 10.1371/journal.pntd.0004299
  212. 212. Jacobs RT, Plattner JJ, Keenan M. Boron-based drugs as antiprotozoals. Current Opinion in Infectious Diseases. 2011;24(6):586-592. DOI: 10.1097/QCO.0b013e32834c630e
  213. 213. Bustamante JM, Craft JM, Crowe BD, Ketchie SA, Tarleton RL. New, combined, and reduced dosing treatment protocols cure Trypanosoma cruzi infection in mice. The Journal of Infectious Diseases. 2014;209(1):150-162. DOI: 10.1093/infdis/jit420
  214. 214. Khare S, Roach SL, Barnes SW, Hoepfner D, Walker JR, Chatterjee AK, et al. Utilizing chemical genomics to identify cytochrome b as a novel drug target for Chagas disease. PLoS Pathogens. 2015;11(7):e1005058. DOI: 10.1371/journal.ppat.1005058

Written By

Luís Gaspar, Terry K. Smith, Nilmar Silvio Moretti, Sergio Schenkman and Anabela Cordeiro-da-Silva

Submitted: 16 May 2017 Reviewed: 06 April 2018 Published: 12 September 2018