Next Article in Journal
Design, Synthesis and Antifungal Evaluation of Novel Pyrylium Salt In Vitro and In Vivo
Next Article in Special Issue
Effects of the Calix[4]arene Derivative Compound OTX008 on High Glucose-Stimulated ARPE-19 Cells: Focus on Galectin-1/TGF-β/EMT Pathway
Previous Article in Journal
Preparation of Polycarbonate-ZnO Nanocomposite Films: Surface Investigation after UV Irradiation
Previous Article in Special Issue
Novel c-Jun N-Terminal Kinase (JNK) Inhibitors with an 11H-Indeno[1,2-b]quinoxalin-11-one Scaffold
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Insights on the Modulation of SIRT5 Activity: A Challenging Balance

Department of Pharmaceutical Sciences, University of Milan, Via L. Mangiagalli 25, 20133 Milano, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2022, 27(14), 4449; https://doi.org/10.3390/molecules27144449
Submission received: 9 June 2022 / Revised: 8 July 2022 / Accepted: 9 July 2022 / Published: 12 July 2022

Abstract

:
SIRT5 is a member of the Sirtuin family, a class of deacetylating enzymes consisting of seven isoforms, involved in the regulation of several processes, including gene expression, metabolism, stress response, and aging. Considering that the anomalous activity of SIRT5 is linked to many pathological conditions, we present herein an overview of the most interesting modulators, with the aim of contributing to further development in this field.

1. Introduction

Class III histone deacetylases (HDACs), commonly known as Sirtuins (SIRTs), are NAD+-dependent enzymes characterized by either histone deacetylase or, more frequently, ribosyl transferase activity [1]. SIRTs are involved in several processes, including transcription, translation, metabolism, and cellular stress/damage. They catalyze different types of chemical reactions, such as demalonylation, desuccinylation, depropionylation, and delipoamidation [2]. Besides their role in gene expression and metabolism, SIRTs are implicated in delaying cellular aging and controlling the normal cell cycle; moreover, they are important mediators of apoptosis inhibition and glucose homeostasis. Interestingly, the downregulation of these enzymes has been associated with metabolic diseases, cancer, and neurodegeneration.
SIRTs comprise seven isoforms, sharing a conserved catalytic core and differing in their N- and C-terminal sequences, which contribute to their specific localization and regulation. SIRT1 has an N-terminal STAC (sirtuin-activating compound) binding domain, SIRT3, 4, and 5 are characterized by a peculiar N-terminal mitochondrial localization sequence (MLS), while SIRT6 and SIRT7 have N- and C-terminal portions contributing to the binding of DNA and chromatin [3].
The different SIRT isoforms catalyze the same deacetylation reaction: NAD+ reacts with the acyclic oxygen of the protein substrate to form a 1′-O-alkylamidate intermediate; then, subsequent hydrolysis provides the deacylated protein and 2′-O-acyl-ADP-ribose. In this reaction, Zn2+ plays a fundamental role as an enzymatic cofactor (Scheme 1) [3].
However, despite the structural similarity among all isoforms, only SIRT5 catalyzes the cleavage of negatively charged acylated substrates. SIRT5 is mainly located in the mitochondria (together with SIRT3 and SIRT4) and is highly expressed in the brain, heart, testis, and lymphoblasts. Its unique structure consists of fourteen α helices and nine β strands, organized in two domains: the Zn2+-binding domain and the Rossmann fold domain (Figure 1) [4]. Moreover, SIRT5 presents two binding pockets, one for the protein substrate and the other for the NAD+ cofactor, both located in the interspace between the Zn2+-binding domain and the Rossmann fold domain. Differently from other sirtuins, SIRT5 has a peculiar biochemistry and distinctive amino acid residues in the active site, namely Ala86, Tyr102, and Arg105. In the catalytic domain, the two non-hydrophobic residues, Tyr102 and Arg105, form hydrogen bonds and ionic-bond interactions with the carboxyl group of the succinyl-lysine substrate [5]. Furthermore, the presence of Ala86 makes the SIRT5 acyl-lysine binding pocket larger than that of other SIRTs.
SIRT5 is highly expressed in the mitochondrial matrix and has a specific mitochondrial localization sequence (MLS); however, it is also present in the cytosol. Recent studies have reported that an alternative splicing of the SIRT5 mRNA leads to two dominant isoforms, SIRT5iso1 and SIRT5iso2, containing the same N-terminal MLS, but differing at the C-terminal portion. While SIRT5iso1 was found only in the cytosol, SIRT5iso2 was also detected in the cytoplasm [5]. Moreover, two additional SIRT5 isoforms (SIRT5iso3 and SIRT5iso4) have been reported in the NCBI database; however, no information is available regarding their expression, localization, or functional properties [6].
SIRT5 catalyzes the deacetylation, demalonylation, desuccinylation, and deglutarylation of lysine, modulating many metabolic enzymes by post-translational modifications (Figure 2). More specifically, it is involved in several cellular metabolic pathways, anti-inflammatory processes, antitumor mechanisms, and in the regulation of the response to oxidative stress [7].
The mitochondrial isoform of SIRT5 regulates ammonia metabolism, fatty acid oxidation, glycolysis, the tricarboxylic acid (TCA) cycle, the electron transport chain (ETC), and apoptosis [8]. Its specific substrates include carbamoyl phosphate synthetase 1 (CPS1) and glutaminase (GLS), which regulate the balance between the urea cycle and mitophagy. Furthermore, SIRT5 demalonylates glyceraldehyde 3-phosphate dehydrogenase (GAPDH) and other enzymes in the glycolytic cascade, leading to increased GAPDH activity. It desuccinylates pyruvate kinase M2 (PKM2) at Lys311 to increase pyruvate kinase activity, thereby promoting the glycolytic flux [3]. In contrast, Xiangyun and co-workers reported that SIRT5 desuccinylates PKM2 at Lys498 to inhibit its activity in tumor cells [9]. Additionally, SIRT5 is specifically involved in the pentose phosphate pathway, which allows the production of ribose-5-phosphate, a precursor for the synthesis of nucleotides. This cycle also promotes the reduction of NADP+ to NADPH, blocking ROS storage in cells and playing a role in the regeneration of the reduced form of GSH (an ROS scavenger). In detail, SIRT5 mediates the deglutarylation and consequent activation of glucose-6-phosphate-1-dehydrogenase (G6PD) in stress conditions to maintain cellular redox homeostasis [8]. Furthermore, SIRT5 regulates 3-hydroxy-3-methylglutaryl-CoA synthase 2 (HMGCS2), enoyl-CoA hydratase α-subunit (ECHA), and malonyl-CoA decarboxylase (MCD). The latter catalyzes the conversion of malonyl-CoA to acetyl-CoA and is implicated in the homeostasis of these metabolites in mitochondria and peroxisomes [10]. SIRT5 also desuccinylates superoxide dismutase [Cu-Zn] (SOD1) and isocitrate dehydrogenase 2 (IDH2) [11].
The cytosolic SIRT5 isoform acts as a glycolytic enzyme and interacts with the 40S and 60S ribosomal subunits during gene translation [5].
Most notably, SIRT5 is also involved in several intracellular signaling pathways, exerting a specific influence in the regulation of protein functions through their post-translational modification (PTM) in physiological and pathological conditions [12,13,14,15,16]. However, the functional role of SIRT5 in regulating many of its proposed targets has yet to be investigated.

2. SIRT5 Modulators

Due to the potential role of SIRT5 as a pharmacological target in cancer, diabetes, cardiovascular diseases, obesity, neurodegenerative disorders, and inflammation, many studies have been undertaken to identify new molecules acting as SIRT5 activators or inhibitors.
Here, we report a comprehensive overview of the SIRT5 modulators that have been described in the literature in the last decade. All selected molecules are classified as activators or inhibitors based on their effects on SIRT5, and further divided into natural derivatives and synthetic compounds. The screening process and the subsequent selection of the relevant literature sources for this review are described in Figure 3 through a PRISMA flow diagram.
The first part of the following two sections is dedicated to natural products (NPs), which represent an invaluable source of structurally unique bioactive compounds, and a paramount source of new drug candidates [17,18,19,20,21]. Semisynthetic and synthetic derivatives are described in the second part of both sections. The origin of the compounds, their biological activities and mode of action, as well as viable opportunities for their future development will be the focus of this work.

2.1. Activators

SIRT5 can be activated by sirtuin-activating compounds (STACs), which induce cellular and physiological effects amplified by downstream signaling pathways. These molecules could be useful in hyperlipidemia and hypercholesterolemia, in the prevention and treatment of type 2 diabetes [22], in the therapeutic approach for Alzheimer’s disease [23], or as medication for rheumatoid arthritis [24] and various tumors [14,25].

2.1.1. Natural Compounds

In a 2006 patent application, Howitz et al. assayed a group of compounds, previously reported as SIRT1 activators, for their ability to activate SIRT5 [22]. Among the detected STACs, several were natural compounds; the most interesting are discussed in the following paragraph. Resveratrol (1) and pinosylvin (2) are natural stilbenes obtained from plants: 1 is commonly found in peanuts, grapes, raspberries, and vegetables and it is commonly used as a dietary supplement, whereas 2 is synthesized in plants during fungal infections, in response to stress and physical damage [26]. Isoliquiritigenin (3) and butein (4) are bioactive chalcone derivatives. Compound 3, found in Glycyrrhiza glabra, is known to be capable of reducing oxidative damage in diabetic patients affected by neuropathy [27], whereas 4, a compound found in several plants, including Toxicodendron vernicifluum, Dahlia spp., and Butea spp., was identified as an antioxidant agent involved in the control of oxidative stress and apoptosis. It was also described as a neuroprotective compound, supporting the correct cerebral activity via an increase in neuronal cell survival and the prevention of H2O2-mediated neurotoxicity [28]. Two flavone derivatives were also identified as STACs, namely fisetin (5), a natural flavonoid found in fruits, vegetables, and nuts, known for its antioxidant and anti-inflammatory properties [29], and quercetin (6), a polar auxin transport inhibitor (ATI), widely distributed in nature [30]. Finally, nordihydroguaiaretic acid (7), a natural antioxidant and anti-inflammatory lignan found in Larrea tridentata, showed activity on SIRT5 [31,32]. Compounds 17 were evaluated for their ability to activate SIRT5 at 200 µM; the results were expressed as the rate of activation with respect to the control (1: 4.95 ± 0.48-fold; 2: 2.71 ± 0.092-fold; 3: 3.93 ± 0.30-fold; 4: 2.19 ± 0.10-fold; 5: 2.61 ± 0.19-fold; 6: 2.18 ± 0.10-fold; 7: 1.91 ± 0.02-fold) [22]. The chemical structures of compounds 17 are reported in Figure 4. To summarize, the reported compounds can be classified into four major groups of polyphenols, namely stilbenes (1, 2), chalcones (3, 4), flavones (5, 6), and lignans (7).
Compound 1 was also studied by Gertz et al., who evaluated its effects on SIRT5 and SIRT3 [33]. The biological analyses revealed a 2.5-fold stimulation of the deacetylase activity of SIRT5 against the fluorophore-modified peptide substrate ‘‘Fluor-de-Lys’’ (FdL) in response to 0.2 µM resveratrol (1). Compound 1 was also shown to inhibit the human SIRT3 and stimulate SIRT1. Unfortunately, 1 is characterized by low solubility, which does not allow full inhibition of SIRT3 and maximum stimulation of SIRT5 to be achieved. The low bioavailability of resveratrol (1) prompted the hypothesis that its metabolites might be responsible for the in vivo effects. The biological analysis of piceatannol (8, Figure 4), a metabolite of 1 showing higher solubility, revealed that the compound stimulates SIRT5 and inhibits SIRT3 (EC50 of 0.07 ± 0.02 mM for SIRT5) [33,34].

2.1.2. Synthetic Compounds

In addition to the natural molecules reported in the previous section, Howitz et al. assayed several synthetic compounds, previously known as SIRT1 activators, on SIRT5 [22]. Most of the tested molecules were stilbene derivatives, including BML-217 (9, Figure 5), which showed the maximum stimulation of SIRT5 (13.6-fold). Interesting results were also obtained for dipyridamol (10, Figure 5; 8.56 ± 0.30-fold), a nucleoside transport and PDE3 inhibitor used in the prevention of blood clot formation [35], and for ZM336372 (11, Figure 5; 0.74 ± 0.05-fold), a c-Raf inhibitor [22,36].
Sauve et al. synthesized several nicotinamide riboside derivatives and assayed them on SIRT5 [37]. The most interesting result was found for 12 (Figure 5), which showed 80% activation at 600 µM [37].
More recently, Hu and co-workers studied the role of SIRT5 in pancreatic ductal adenocarcinoma (PDAC) and developed a small-molecule SIRT5 activator, MC3138 (13, Figure 5) [38]. The compound was tested on human PDAC cell lines, organoids, and PDX tumors. The results of the biochemical enzymatic assays indicated that 13 exhibited the selective activation of SIRT5 over SIRT1/3. The cell viability experiments showed that 13 reduced PDAC cell viability, with IC50 values ranging from 25.4 mmol/L to 236.9 mmol/L. The toxicity assays demonstrated that 13 combined with gemcitabine could be a safe and effective therapeutic option for PDAC characterized by low SIRT5 expression [38].
In a computational study, Schlicker et al. performed docking screening on a library of small molecules, using the crystal structures of human SIRT2/3/5/6 as models [39]. The virtual simulations suggested that CSC9 (14), CSC33 (15), and CSC38 (16) (Figure 5) could act as activators of SIRT5 [39].

2.2. Inhibitors

SIRT5 inhibitors (sirtuin-inhibiting compounds, STICs) have been studied for their potential role in the treatment of metabolic disorders [6,40], neurodegenerative pathologies [23,41,42], cardiovascular diseases [42,43], and cancer [3,44].

2.2.1. Natural Compounds

Nicotinamide (NAM, 17, Figure 6) represents one of the major precursors for NAD biosynthesis. As a product of the reactions catalyzed by SIRTs, it acts as an endogenous non-competitive inhibitor of these enzymes, with an IC50 = 150 μM against SIRT5 [45].
In 2015, Huang et al. performed a screening on an NP database, discovering interesting SIRT5-inhibitory activity for oleanolic acid (18) and echinocystic acid (19) (Figure 6), with IC50 values of 70 and 40 μM, respectively [46]. Compound 18 is widely distributed in food (olive oil, garlic, etc.) and plants (Phytolacca americana, Syzygium spp., Rosa woodsi, etc.), and it exhibits hepatoprotective, antitumor, and antiviral properties. Compound 19 (Figure 6) is a triterpenoid derivative found in several plants, such as Codonopsis lanceolata, Cucurbita foetidissima, and Eclipta alba, and is known to have anti-inflammatory, antiviral, and anticancer activities [46].
Guetschow et al. identified several natural SIRT5 inhibitors via a high-throughput screening of a Prestwick Chemical library, comprising 1280 approved drugs [47]. Eight of them were selected as promising inhibitors of SIRT5, among which four were natural compounds: antymicin (20, a commercial piscicide used in catfish production), thyroxine (21, a thyroid medicine administered in the treatment of goiter and thyroid nodules), anthralin (22, a topical drug for the treatment of local diseases, such as psoriasis), and methacycline (23, a derivative of tetracycline used against Gram-positive, Gram-negative, and L-form bacteria) [47,48]. The IC50 values against SIRT5 were calculated as follows: 20, IC50 = 90 μM; 21, IC50 = 2.2 μM; 22, IC50 = 0.1 μM; 23, IC50 = 3.6 μM [48]. The structures of compounds 2023 are reported in Figure 6.

2.2.2. Synthetic Compounds

Over the years, several research groups have based the design of new SIRT5 inhibitors on the discovery that this isoform preferentially catalyzes the hydrolysis of succinyl and malonyl groups, rather than acetyl groups, from lysine residues. Most notably, SIRT5 is the only SIRT that exhibits this unique preference. Therefore, this characteristic has been conveniently exploited to develop selective inhibitors of SIRT5 [1].
In 2012, He and co-workers reported for the first time the synthesis of a histone H3 lysine 9 (H3K9) thiosuccinyl peptide (H3K9Tsu, 24, Figure 7) as a selective SIRT5 inhibitor [49]. They designed the molecule as a mechanism-based inhibitor, potentially capable of blocking the deacetylase activity of the enzyme by the formation of a stalled covalent intermediate. Compound 24 exhibited an IC50 of 5 μM against SIRT5 and was inactive (IC50 >100 μM) on other isoforms; the selectivity was achieved based on the inability of other SIRTs to recognize malonyl and succinyl lysine peptides. However, the compound was not active in whole-cell assays due to its low permeability [49]. Encouraged by these results, Lin reported the discovery of the first series of thiourea derivatives as potent inhibitors of SIRT5 [50]. The most interesting compounds, JH-I5-2 alias 25, 26, and 27, exhibited IC50 values of 0.89, 0.45, and 12 μM, respectively [50]. These peptides became the starting point for a new class of analogues containing the thiourea moiety. In this context, Negròn Abril et al. developed cell-permeable SIRT5-selective inhibitors [51]. Among the synthesized derivatives, DK1-04 (28, Figure 7) showed the strongest inhibition, with an IC50 of 0.34 µM (no SIRT1-3,6 inhibition at 83.3 µM). Because 28 contained a free carboxylic acid, which could hinder cellular permeability, the function was protected with either an acetomethoxyl or an ethyl ester, which could be easily hydrolyzed in cells to release the active form of the compound. The prodrugs were also tested in 2D proliferation assays on MCF7 and MDA-MB-231 breast cancer cells: DK1-04e (29) showed the strongest inhibition of cell growth. The researchers also performed in vivo studies using a mouse model of breast cancer, MMTV-PyMT. The treatment with 29 (50 mg/kg, daily for 3 weeks) significantly reduced the tumor size and weight. Importantly, 29 did not cause any apparent toxicity or significant weight loss in mice. These results suggest that SIRT5 inhibitors have promising potential as innovative treatment options for breast cancer [51].
Many other research groups investigated thiourea derivatives, albeit with more modest success. In 2016, Liu and co-workers synthesized peptide 30 (Figure 7) [52], characterized by a central Nε-carboxyethyl-thiocarbamoyl-lysine residue. Compound 30 showed an IC50 of 7.6 ± 1.5 μM against SIRT5 and very weak inhibitory activity against SIRT1/3/6 (IC50 > 1000 μM). Unfortunately, its inhibition of SIRT2 was found to be only around 13-fold weaker than that of SIRT5 (IC50 = 96.4 ± 18.5 μM). The same research group also synthesized some cyclic derivatives of 30, which are reported in the next section [52]. Recently, Yang et al. prepared a series of 3-thioureidopropanoic acid derivatives mimicking glutaryl-lysine substrates [53]; among them, 31 (Figure 7) showed promising inhibitory activity and selectivity for SIRT5 (IC50 = 3.0 μM, SIRT1-3,6 IC50 > 600 μM) [53].
Among a series of compounds bearing the thiourea warhead reported by Rajabi et al., 32 (Figure 8) exhibited the most potent and selective SIRT5-inhibitory activity (IC50 = 0.11 μM) [54]. This peptide became the lead compound for a series of optimization campaigns, carried out by different research groups. In 2022, Rajabi and co-workers investigated isostere and prodrug derivatives of 32, modified at the carboxylic function [55]. In a BioRxiv preprint, Bolding et al. reported the study of fluorosulfate analogues, with potential activity in vivo [56]. From the same class, Yan and co-workers identified a prodrug, NRD167 (33, Figure 8), which was able to block the proliferation of the SIRT5-dependent cell lines SKM-1 and OCI-AML2 (IC50 of 5 and 8 μmol/L, respectively), and to induce >80% apoptosis (at 5 and 10 μmol/L, respectively) [57]. Zang et al. identified the small peptide 34 (Figure 8), endowed with modest activity on SIRT5 (IC50 = 5 μM) and high selectivity vs. SIRT1 and 6 [58]. In the same year, He and co-workers discovered weaker activity for the analogue 35 (Figure 8, IC50 = 92.1 ± 3.5 μM), while studying the mechanism of action of several pan-SIRT1-3 derivatives bearing an l-2-amino-7-carboxamidoheptanoic acid (l-ACAH) moiety [59].
In 2015, Polletta and co-workers investigated the effect of SIRT5 on ammonia detoxification [60]. Their results showed that ammonia production increases in SIRT5-silenced and decreases in SIRT5-overexpressing cells. They also obtained the same ammonia increase when using a new specific inhibitor of SIRT5, MC3482 (36, Figure 7). Compound 36 inhibited the desuccinylase activity of SIRT5 (42% inhibition at 50 μM), affecting glutamine metabolism by interaction with the GLS catalytic domain. Although it was quite selective and did not inhibit SIRT1/3, it exhibited low in vitro potency [60].
A campaign to design analogues of CPS1, a substrate of SIRT5, was undertaken by Roessler and co-workers [61]. Among their peptide derivatives, the most promising was 37 (Figure 8), which showed a Ki value of 4.3 μM against SIRT5 and good results in the selectivity assays (Ki > 50 μM against SIRT1/2/3) [61]. Moreover, 38 (Figure 8) exhibited significant inhibition, with a Ki of 38.1 ± 0.63 μM [61]. In a similar effort, Kalbas et al. developed several compounds related to CPS1, substituted at position 3 of the succinyl moiety; among them, 39 (Figure 8) showed very potent SIRT5 inhibition (IC50 = 15.4 nM) [62].
Peptide macrocyclization is an efficient approach to confer upon a linear peptide enhanced metabolic stability and cell permeability. Therefore, the synthesis of cyclic peptides from linear sequences could be an interesting technique for the development of innovative SIRT5 inhibitors.
On these premises, Liu et al. synthesized the cyclic derivatives 40 and 41 (Figure 9) of the linear peptide 30 (Figure 7) [52]. These compounds exhibited comparable inhibitory potency to that of the parent peptide (40, IC50 = 6.0 ± 3.0 μM; 41, IC50 = 7.5 ± 4.0 μM), suggesting that the macrocyclic bridging units in 40 and 41 were unable to constrain the peptide backbone into a bioactive conformation, or interfered with the binding of the compounds to the SIRT5 active site. Interestingly, the same macrocyclic bridging units were able to significantly enhance the potency of a similar linear peptide against SIRT1/2/3/6 [52,63,64].
Later, Jiang and co-workers prepared several cyclic tripeptides endowed with higher activity with respect to 40 and 41 [65]. Among them, 42 (Figure 9) exhibited the highest activity and selectivity vs. SIRT5 (IC50(SIRT5) = 2.2 ± 0.89 μM; IC50(SIRT1) = 254.2 ± 32.4 μM; IC50(SIRT2) = 131.3 ± 46.5 μM; IC50(SIRT3) > 450 μM; IC50(SIRT6) > 1000 μM) [65].
Peptide derivatives are usually characterized by low biostability and scarce membrane permeability; therefore, the identification of small molecules is crucial to overcome these limits.
Liu and co-workers synthesized a library of (E)-2-cyano-N-phenyl-3-(5-phenylfuran-2-yl)acrylamide derivatives, among which 43 (Figure 10) emerged as the most potent inhibitor of SIRT5, with an IC50 of 5.59 ± 0.75 μM [66]. Further biochemical studies revealed that 43 likely acts by competing with the succinyl-lysine substrate for the ligand-binding site of SIRT5. The compound also exhibited promising selectivity for SIRT5 over SIRT2/6 [66].
While verifying the cross-reactivity of a series of SIRT inhibitors towards SIRT5, Suenkel et al. found that GW5074 (44, Figure 10), a known SIRT2 blocker, showed interesting activity against SIRT5 [67]. The effect of 44 on SIRT5 deacetylation was comparable to that exerted on SIRT2 (more than 40% inhibition at 12.5 μM). However, 44 was also active against kinases; hence, it was deemed unsuitable for further in vivo studies. Nonetheless, it remains a promising starting point for the development of specific SIRT5 inhibitors [67].
Suramin (45, Figure 10), a drug approved for the treatment of human sleeping sickness caused by trypanosomes, was identified by Schuetz A. et al. as an inhibitor of SIRT5, with an IC50 value of 22 μM [68]. Two crystal structures of SIRT5, one in complex with ADP-ribose and the other with suramin (45), were reported by the authors. The analysis of the structures revealed that 45 acts as a linker molecule, resulting in the dimerization of SIRT5. This finding may lead to the development of a new class of inhibitors that not only bind specifically to the active site of the enzyme, but also function as linker molecules, thus limiting enzyme mobility and accessibility. Unfortunately, the fact that suramin (45) also targets the NAD+-binding region of SIRT5 makes it non-selective for similar binding pockets. Indeed, 45 has been reported to inhibit various other NAD+/NADP+-dependent enzymes, such as the lactate dehydrogenase from Dirofilaria immitis and Onchocerca volvulus, and the glyceraldehyde-3-phosphate dehydrogenase from Trypanosoma brucei. This highlights that the most efficient strategy to develop selective SIRT inhibitors is to target the substrate-binding site, and not the NAD+-binding pocket, in order to avoid secondary effects [68].
Among the compounds identified by Guetschow and co-workers via the high-throughput screening approach described in the previous section, four synthetic approved drugs exhibited promising inhibitory activity vs. SIRT5 [47]. Probucol (46), an anti-hyperlipidemic drug, showed an IC50 of 1.6 μM; fulvestrant (47), an estrogen receptor antagonist used in hormonal therapy, had an IC50 of 2.6 μM; balsalazide (48), a drug used for the treatment of inflammatory bowel disease, exhibited an IC50 of 3.9 μM; finally, closantel (49), a veterinary drug with potent antiparasitic activity, displayed an IC50 of 2.7 μM [47,48]. The chemical structures of compounds 4649 are reported in Figure 10. Inspired by this study, Glas et al. selected balsalazide (48) as the starting point for the development of new synthetic derivatives [48]. Unfortunately, none of the analogues showed improved inhibitory activity with respect to the parent compound. In detail, changes to the N-aroyl-β-alanine side chain disrupted the activity, while the introduction of a truncated salicylic acid moiety minimally altered the potency. The most interesting derivative was 50 (Figure 10), which showed a minimal loss of activity, demonstrating that modifications on the salicylic acid portion are tolerated, to some extent (SIRT5 inhibition = 73% at 50 μM). Interestingly, the biological assays proved that 48 and its derivatives were selective for SIRT5 over SIRT1/2/3. Unfortunately, 48 cannot be considered as an optimal candidate for the inhibition of SIRT5, due to its minimal absorption from the gut, poor water solubility, and instability to enzymatic degradation by the intestinal microflora (reductive cleavage of the azo moiety) [48].
As previously discussed, Schlicker and colleagues performed docking screening on a library of small molecules [39]. Several potential SIRT ligands were identified, but further activity tests were conducted only against SIRT2. Docking studies suggested that CSC1 (51), CSC14 (52), and CSC21 (53) could act as inhibitors of different isoforms of SIRT, including SIRT5 (Figure 11) [39].
The research group of Maurer and colleagues worked on the identification of small molecules as SIRT5 inhibitors; the compounds showed good potencies, but very poor selectivity vs. other SIRT isoforms [69]. In detail, two naphthol derivatives, cambinol (54) and sirtinol (55), already known for their anti-inflammatory properties [70,71], were proven to possess non-selective activity against SIRT5 at submicromolar concentrations (Figure 11) [69]. Moreover, they prepared several potent thiobarbiturate derivatives, which were also active against SIRT1/2. However, they were shown to be selective over SIRT3, an isoform that is co-localized with SIRT5. The thiobarbiturate derivative 56 (Figure 11) exhibited the highest inhibitory activity against SIRT5 (IC50(SIRT1) = 5.3 ± 0.7 μM, IC50(SIRT2) = 9.7 ± 1.6 μM, SIRT3: 41% inhibition at 50 μM, IC50(SIRT5) = 2.3 ± 0.2 μM) [69]. Other potent SIRT5 inhibitors lacking the desired selectivity were synthesized by Han et al., based on the 8-mercapto-3,7-dihydro-1H-purine-2,6-dione scaffold [72]. The structure of one of the most potent derivatives (57) is reported in Figure 11 (IC50(SIRT1) = 0.12 ± 0.01 μM, IC50(SIRT2) = 1.19 ± 0.06 μM, IC50(SIRT3) = 0.54 ± 0.05 μM, IC50(SIRT5) = 0.39 ± 0.03, IC50(SIRT6) = 128.7 ± 16.1 μM) [72].
Several new 9-substituted norharmane derivatives were studied by Yang and co-workers as SIRT5 inhibitors; the most active candidate, 58 (Figure 11), showed 35 and 52% inhibition at 30 μM and 100 μM, respectively [73]. This series was subsequently developed into a library of derivatives characterized by different linker bridges and phenyl substituents, which exhibited significant activities [74]. During the development of assay platforms based on fluorogenic substrates to identify SIRT inhibitors, the same group discovered that TW-37 (59, Figure 11), a Bcl-2 inhibitor, was capable of inhibiting SIRT5 in the low micromolar range (IC50 = 6 μM) [75].

3. Conclusions

Because SIRTs play essential roles in cell signaling pathways, gene translation, metabolism, and oxidative stress control, they are involved in several pathological conditions, including cancer, diabetes, and neurodegenerative and cardiovascular diseases. Specifically, SIRT5 is an attractive enzyme that not only catalyzes deacetylation reactions, but also exhibits strong demalonylase, desuccinylase, and deglutarylase activities. Consequently, SIRT5 is considered to be a promising molecular target for the treatment of several human diseases.
To date, the number of SIRT5 inhibitors reported in the literature far exceeds that of SIRT5 activators. Moreover, several known compounds and drugs have been identified as SIRT5 modulators by various screening approaches; however, some of them are non-selective and certain ones are considered pan-SIRT modulators (e.g., compounds 1, 17, 5456, etc.).
Overall, despite the promising results achieved in SIRT research and the potential of recent perspectives concerning the role of SIRT5 as a pro-viral factor [76] and as a catalyst in peptide self-assembly [77], it is necessary to consider the issue of selectivity. This is especially crucial because some SIRTs are co-present in the same cellular compartment (for example, SIRT5 and SIRT3 are both localized in the mitochondrial matrix). Furthermore, SIRT5 has pleiotropic roles in tumorigenesis, acting as a tumor suppressor or an oncogene via post-translational modifications, depending on cell conditions [6,44,78]. Hence, the balance between its activation and inhibition should be carefully considered in the development of a SIRT5 modulator.
In conclusion, SIRT5 is an interesting molecular target, and the regulation of its activity has many potential therapeutic applications for a variety of medical conditions. However, additional studies are crucial to identify the molecular determinants to achieve a selective effect, and to verify the feasibility of future pharmaceutical applications.

Author Contributions

Conceptualization, S.V. and A.G.; methodology, A.G. and M.M. Validation, M.M., G.C. and A.G.; formal analysis, S.V. and G.C.; investigation, A.G.; resources, M.M.; data curation, G.C. and M.M.; writing—original draft preparation, A.G., G.C. and M.M.; writing—review and editing, M.M. and G.C.; visualization, F.M.; supervision, S.V. and A.G.; project administration, A.G.; funding acquisition, F.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors acknowledge Servier Medical Art (https://smart.servier.com, accessed on 7 July 2022) for providing the template images (cell and mitochondrion) used in Figure 2.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, L.; Ma, X.; He, Y.; Yuan, C.; Chen, Q.; Li, G.; Chen, X. Sirtuin 5: A review of structure, known inhibitors and clues for developing new inhibitors. Sci. China Life Sci. 2016, 60, 249–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Imai, S.-I.; Armstrong, C.M.; Kaeberlein, M.; Guarente, L. Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature 2000, 403, 795–800. [Google Scholar] [CrossRef]
  3. Wang, Y.; Chen, H.; Zha, X. Overview of SIRT5 as a potential therapeutic target: Structure, function and inhibitors. Eur. J. Med. Chem. 2022, 236, 114363. [Google Scholar] [CrossRef] [PubMed]
  4. Hang, T.; Chen, W.; Wu, M.; Zhan, L.; Wang, C.; Jia, N.; Zhang, X.; Zang, J. Structural insights into the molecular mechanism underlying Sirt5-catalyzed desuccinylation of histone peptides. Biochem. J. 2019, 476, 211–223. [Google Scholar] [CrossRef]
  5. Bringman-Rodenbarger, L.R.; Guo, A.H.; Lyssiotis, C.A.; Lombard, D.B. Emerging Roles for SIRT5 in Metabolism and Cancer. Antioxid. Redox Signal. 2018, 28, 677–690. [Google Scholar] [CrossRef]
  6. Kumar, S.; Lombard, D.B. Functions of the sirtuin deacylase SIRT5 in normal physiology and pathobiology. Crit. Rev. Biochem. Mol. Biol. 2018, 53, 311–334. [Google Scholar] [CrossRef]
  7. Nishida, Y.; Rardin, M.J.; Carrico, C.; He, W.; Sahu, A.K.; Gut, P.; Najjar, R.; Fitch, M.; Hellerstein, M.; Gibson, B.W.; et al. SIRT5 Regulates both Cytosolic and Mitochondrial Protein Malonylation with Glycolysis as a Major Target. Mol. Cell 2015, 59, 321–332. [Google Scholar] [CrossRef] [Green Version]
  8. Yang, X.; Wang, Z.; Li, X.; Liu, B.; Liu, M.; Liu, L.; Chen, S.; Ren, M.; Wang, Y.; Yu, M.; et al. SHMT2 Desuccinylation by SIRT5 Drives Cancer Cell Proliferation. Cancer Res. 2018, 78, 372–386. [Google Scholar] [CrossRef] [Green Version]
  9. Xiangyun, Y.; Xiaomin, N.; Linping, G.; Yunhua, X.; Ziming, L.; Yongfeng, Y.; Zhiwei, C.; Shun, L.; Xiangyun, Y.; Xiaomin, N.; et al. Desuccinylation of pyruvate kinase M2 by SIRT5 contributes to antioxidant response and tumor growth. Oncotarget 2016, 8, 6984–6993. [Google Scholar] [CrossRef] [Green Version]
  10. Colak, G.; Pougovkina, O.; Dai, L.; Tan, M.; Brinke, H.T.; Huang, H.; Cheng, Z.; Park, J.; Wan, X.; Liu, X.; et al. Proteomic and Biochemical Studies of Lysine Malonylation Suggest Its Malonic Aciduria-associated Regulatory Role in Mitochondrial Function and Fatty Acid Oxidation. Mol. Cell. Proteom. 2015, 14, 3056–3071. [Google Scholar] [CrossRef] [Green Version]
  11. Lin, Z.-F.; Xu, H.-B.; Wang, J.-Y.; Lin, Q.; Ruan, Z.; Liu, F.-B.; Jin, W.; Huang, H.-H.; Chen, X. SIRT5 desuccinylates and activates SOD1 to eliminate ROS. Biochem. Biophys. Res. Commun. 2013, 441, 191–195. [Google Scholar] [CrossRef] [PubMed]
  12. Guan, J.; Jiang, X.; Gai, J.; Sun, X.; Zhao, J.; Li, J.; Li, Y.; Cheng, M.; Du, T.; Fu, L.; et al. Sirtuin 5 regulates the proliferation, invasion and migration of prostate cancer cells through acetyl-CoA acetyltransferase 1. J. Cell. Mol. Med. 2020, 24, 14039–14049. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, X.; Rong, F.; Tang, J.; Zhu, C.; Chen, X.; Jia, S.; Wang, Z.; Sun, X.; Deng, H.; Zha, H.; et al. Repression of p53 function by SIRT5-mediated desuccinylation at Lysine 120 in response to DNA damage. Cell Death Differ. 2021, 29, 722–736. [Google Scholar] [CrossRef] [PubMed]
  14. Choi, S.Y.; Jeon, J.M.; Na, A.Y.; Kwon, O.K.; Bang, I.H.; Ha, Y.-S.; Bae, E.J.; Park, B.-H.; Lee, E.H.; Kwon, T.G.; et al. SIRT5 Directly Inhibits the PI3K/AKT Pathway in Prostate Cancer Cell Lines. Cancer Genom. Proteom. 2021, 19, 50–59. [Google Scholar] [CrossRef]
  15. Tang, S.-J.; Yang, J.-B. LncRNA SNHG14 aggravates invasion and migration as ceRNA via regulating miR-656-3p/SIRT5 pathway in hepatocellular carcinoma. Mol. Cell. Biochem. 2020, 473, 143–153. [Google Scholar] [CrossRef]
  16. Shang, B.; Xu, T.; Hu, N.; Mao, Y.; Du, X. Circ-Klhl8 overexpression increased the therapeutic effect of EPCs in diabetic wound healing via the miR-212-3p/SIRT5 axis. J. Diabetes Complicat. 2021, 35, 108020. [Google Scholar] [CrossRef]
  17. Atanasov, A.G.; Zotchev, S.B.; Dirsch, V.M.; Orhan, I.E.; Banach, M.; Rollinger, J.M.; Barreca, D.; Weckwerth, W.; Bauer, R.; Bayer, E.A.; et al. Natural products in drug discovery: Advances and opportunities. Nat. Rev. Drug Discov. 2021, 20, 200–216. [Google Scholar] [CrossRef]
  18. Huang, M.; Lu, J.-J.; Ding, J. Natural Products in Cancer Therapy: Past, Present and Future. Nat. Prod. Bioprospecting 2021, 11, 5–13. [Google Scholar] [CrossRef]
  19. Panda, S.S.; Jhanji, N. Natural Products as Potential Anti-Alzheimer Agents. Curr. Med. Chem. 2020, 27, 5887–5917. [Google Scholar] [CrossRef]
  20. Cazzaniga, G.; Mori, M.; Chiarelli, L.R.; Gelain, A.; Meneghetti, F.; Villa, S. Natural products against key Mycobacterium tuberculosis enzymatic targets: Emerging opportunities for drug discovery. Eur. J. Med. Chem. 2021, 224, 113732. [Google Scholar] [CrossRef]
  21. Aswad, M.; Rayan, M.; Abu-Lafi, S.; Falah, M.; Raiyn, J.; Abdallah, Z.; Rayan, A. Nature is the best source of anti-inflammatory drugs: Indexing natural products for their anti-inflammatory bioactivity. Agents Actions 2017, 67, 67–75. [Google Scholar] [CrossRef] [PubMed]
  22. Howitz, K.T.; Zipkin, R.E. Compositions and Methods for Selectively Activating Human Sirtuins. U.S. Patent Application No. 11/166,892, 19 January 2006. [Google Scholar]
  23. Wu, S.; Wei, Y.; Li, J.; Bai, Y.; Yin, P.; Wang, S. SIRT5 Represses Neurotrophic Pathways and Aβ Production in Alzheimer’s Disease by Targeting Autophagy. ACS Chem. Neurosci. 2021, 12, 4428–4437. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, N.; Zhang, H.; Law, B.Y.K.; Dias, I.D.S.R.; Qiu, C.L.; Zeng, W.; Pan, H.D.; Chen, J.Y.; Bai, Y.F.; Lv, J.; et al. Sirtuin 5 deficiency increases disease severity in rats with adjuvant-induced arthritis. Cell. Mol. Immunol. 2020, 17, 1190–1192. [Google Scholar] [CrossRef] [PubMed]
  25. Yihan, L.; Xiaojing, W.; Ao, L.; Chuanjie, Z.; Haofei, W.; Yan, S.; Hongchao, H. SIRT5 functions as a tumor suppressor in renal cell carcinoma by reversing the Warburg effect. J. Transl. Med. 2021, 19, 521. [Google Scholar] [CrossRef] [PubMed]
  26. Lee, S.K.; Lee, H.J.; Min, H.Y.; Park, E.J.; Lee, K.M.; Ahn, Y.H.; Cho, Y.J.; Pyee, J.H. Antibacterial and antifungal activity of pinosylvin, a constituent of pine. Fitoterapia 2005, 76, 258–260. [Google Scholar] [CrossRef]
  27. Yerra, V.G.; Kalvala, A.K.; Kumar, A. Isoliquiritigenin reduces oxidative damage and alleviates mitochondrial impairment by SIRT1 activation in experimental diabetic neuropathy. J. Nutr. Biochem. 2017, 47, 41–52. [Google Scholar] [CrossRef]
  28. Wang, Y.; Chan, F.L.; Chen, S.; Leung, L.K. The plant polyphenol butein inhibits testosterone-induced proliferation in breast cancer cells expressing aromatase. Life Sci. 2005, 77, 39–51. [Google Scholar] [CrossRef]
  29. Gupta, S.C.; Tyagi, A.K.; Deshmukh-Taskar, P.; Hinojosa, M.; Prasad, S.; Aggarwal, B.B. Downregulation of tumor necrosis factor and other proinflammatory biomarkers by polyphenols. Arch. Biochem. Biophys. 2014, 559, 91–99. [Google Scholar] [CrossRef]
  30. Fischer, C.; Speth, V.; Fleig-Eberenz, S.; Neuhaus, G. Induction of Zygotic Polyembryos in Wheat: Influence of Auxin Polar Transport. Plant Cell 1997, 9, 1767–1780. [Google Scholar] [CrossRef]
  31. Wang, L.; Li, L.; Quan, M.-Y.; Wang, D.; Jia, Z.; Li, Z.-F.; Li, B.; Guo, L.; Tan, G.-J. Nordihydroguaiaretic acid can suppress progression of experimental autoimmune encephalomyelitis. IUBMB Life 2018, 70, 432–436. [Google Scholar] [CrossRef]
  32. Gilbert, N.C.; Gerstmeier, J.; Schexnaydre, E.E.; Börner, F.; Garscha, U.; Neau, D.B.; Werz, O.; Newcomer, M.E. Structural and mechanistic insights into 5-lipoxygenase inhibition by natural products. Nat. Chem. Biol. 2020, 16, 783–790. [Google Scholar] [CrossRef]
  33. Gertz, M.; Nguyen, G.T.T.; Fischer, F.; Suenkel, B.; Schlicker, C.; Fränzel, B.; Tomaschewski, J.; Aladini, F.; Becker, C.; Wolters, D.; et al. A Molecular Mechanism for Direct Sirtuin Activation by Resveratrol. PLoS ONE 2012, 7, e49761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Mayack, B.K.; Sippl, W.; Ntie-Kang, F. Natural Products as Modulators of Sirtuins. Molecules 2020, 25, 3287. [Google Scholar] [CrossRef]
  35. Brown, D.G.; Wilkerson, E.C.; Love, W.E. A review of traditional and novel oral anticoagulant and antiplatelet therapy for dermatologists and dermatologic surgeons. J. Am. Acad. Dermatol. 2015, 72, 524–534. [Google Scholar] [CrossRef] [PubMed]
  36. Van Gompel, J.J.; Kunnimalaiyaan, M.; Holen, K.; Chen, H. ZM336372, a Raf-1 activator, suppresses growth and neuroendocrine hormone levels in carcinoid tumor cells. Mol. Cancer Ther. 2005, 4, 910–917. [Google Scholar] [CrossRef] [Green Version]
  37. Sauve, A.; Cen, Y. Activation and activators of SIRT5 2011. U.S. Patent Applicatio No. 20120329748A1, 27 December 2012. [Google Scholar]
  38. Hu, T.; Shukla, S.K.; Vernucci, E.; He, C.; Wang, D.; King, R.J.; Jha, K.; Siddhanta, K.; Mullen, N.J.; Attri, K.S.; et al. Metabolic Rewiring by Loss of Sirt5 Promotes Kras-Induced Pancreatic Cancer Progression. Gastroenterology 2021, 161, 1584–1600. [Google Scholar] [CrossRef]
  39. Schlicker, C.; Boanca, G.; Lakshminarasimhan, M.; Steegborn, C. Structure-based development of novel sirtuin inhibitors. Aging 2011, 3, 852–872. [Google Scholar] [CrossRef] [Green Version]
  40. Elkhwanky, M.-S.; Hakkola, J. Extranuclear Sirtuins and Metabolic Stress. Antioxid. Redox Signal. 2018, 28, 662–676. [Google Scholar] [CrossRef] [Green Version]
  41. Yeong, K.Y.; Berdigaliyev, N.; Chang, Y. Sirtuins and Their Implications in Neurodegenerative Diseases from a Drug Discovery Perspective. ACS Chem. Neurosci. 2020, 11, 4073–4091. [Google Scholar] [CrossRef]
  42. Gomes, P.; Leal, H.; Mendes, A.F.; Reis, F.; Cavadas, C. Dichotomous Sirtuins: Implications for Drug Discovery in Neurodegenerative and Cardiometabolic Diseases. Trends Pharmacol. Sci. 2019, 40, 1021–1039. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, Y.; He, J.; Liao, M.; Hu, M.; Li, W.; Ouyang, H.; Wang, X.; Ye, T.; Zhang, Y.; Ouyang, L. An overview of Sirtuins as potential therapeutic target: Structure, function and modulators. Eur. J. Med. Chem. 2019, 161, 48–77. [Google Scholar] [CrossRef] [PubMed]
  44. Maiese, K. (Ed.) Sirtuin Biology in Cancer and Metabolic Disease; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  45. Fischer, F.; Gertz, M.; Suenkel, B.; Lakshminarasimhan, M.; Schutkowski, M.; Steegborn, C. Sirt5 Deacylation Activities Show Differential Sensitivities to Nicotinamide Inhibition. PLoS ONE 2012, 7, e45098. [Google Scholar] [CrossRef] [PubMed]
  46. Huang, W.-F.; You, L.; Zhu, H.; Liao, S.-G.; Li, Y.-J.; Liu, T.; Wang, A.-M.; He, B.; Lan, Y.-Y. Screening of SIRT5 inhibitors from natural products. Zhongguo Xinyao Zazhi 2015, 24, 2724–2728. [Google Scholar]
  47. Guetschow, E.D.; Kumar, S.; Lombard, D.B.; Kennedy, R.T. Identification of sirtuin 5 inhibitors by ultrafast microchip electrophoresis using nanoliter volume samples. Anal. Bioanal. Chem. 2015, 408, 721–731. [Google Scholar] [CrossRef] [PubMed]
  48. Glas, C.; Dietschreit, J.C.B.; Wössner, N.; Urban, L.; Ghazy, E.; Sippl, W.; Jung, M.; Ochsenfeld, C.; Bracher, F. Identification of the subtype-selective Sirt5 inhibitor balsalazide through systematic SAR analysis and rationalization via theoretical investigations. Eur. J. Med. Chem. 2020, 206, 112676. [Google Scholar] [CrossRef]
  49. He, B.; Du, J.; Lin, H. Thiosuccinyl Peptides as Sirt5-Specific Inhibitors. J. Am. Chem. Soc. 2012, 134, 1922–1925. [Google Scholar] [CrossRef] [Green Version]
  50. Lin, H. Thiourea compounds and their use as inhibitors of SIRT2 or SIRT5 2014. U.S. Patent No. 10,556,878, 11 February 2020. [Google Scholar]
  51. Abril, Y.L.N.; Fernandez, I.R.; Hong, J.Y.; Chiang, Y.-L.; Kutateladze, D.A.; Zhao, Q.; Yang, M.; Hu, J.; Sadhukhan, S.; Li, B.; et al. Pharmacological and genetic perturbation establish SIRT5 as a promising target in breast cancer. Oncogene 2021, 40, 1644–1658. [Google Scholar] [CrossRef]
  52. Liu, J.; Huang, Y.; Zheng, W. A Selective Cyclic Peptidic Human SIRT5 Inhibitor. Molecules 2016, 21, 1217. [Google Scholar] [CrossRef] [Green Version]
  53. Yang, F.; Su, H.; Deng, J.; Mou, L.; Wang, H.; Li, R.; Dai, Q.-Q.; Yan, Y.-H.; Qian, S.; Wang, Z.; et al. Discovery of new human Sirtuin 5 inhibitors by mimicking glutaryl-lysine substrates. Eur. J. Med. Chem. 2021, 225, 113803. [Google Scholar] [CrossRef]
  54. Rajabi, N.; Auth, M.; Troelsen, K.R.; Pannek, M.; Bhatt, D.P.; Fontenas, M.; Hirschey, M.D.; Steegborn, C.; Madsen, A.S.; Olsen, C.A. Mechanism-Based Inhibitors of the Human Sirtuin 5 Deacylase: Structure-Activity Relationship, Biostructural, and Kinetic Insight. Angew. Chem. Int. Ed. 2017, 56, 14836–14841. [Google Scholar] [CrossRef] [Green Version]
  55. Rajabi, N.; Hansen, T.N.; Nielsen, A.L.; Nguyen, H.T.; Bæk, M.; Bolding, J.E.; Bahlke, O.; Petersen, S.E.G.; Bartling, C.R.O.; Strømgaard, K.; et al. Investigation of Carboxylic Acid Isosteres and Prodrugs for Inhibition of the Human SIRT5 Lysine Deacylase Enzyme. Angew. Chem. Int. Ed. 2022, 61, e202115805. [Google Scholar] [CrossRef] [PubMed]
  56. Bolding, J.E.; Martín-Gago, P.; Rajabi, N.; Gamon, L.F.; Hansen, T.N.; Davies, M.J.; Olsen, C.A. Aryl Fluorosulfate-Based Inhibitors that Covalently Target the SIRT5 Lysine Deacylase. bioRxiv 2022. [Google Scholar] [CrossRef]
  57. Yan, D.; Franzini, A.; Pomicter, A.D.; Halverson, B.J.; Antelope, O.; Mason, C.C.; Ahmann, J.M.; Senina, A.V.; Vellore, N.A.; Jones, C.L.; et al. SIRT5 Is a Druggable Metabolic Vulnerability in Acute Myeloid Leukemia. Blood Cancer Discov. 2021, 2, 266–287. [Google Scholar] [CrossRef] [PubMed]
  58. Zang, W.; Hao, Y.; Wang, Z.; Zheng, W. Novel thiourea-based sirtuin inhibitory warheads. Bioorganic Med. Chem. Lett. 2015, 25, 3319–3324. [Google Scholar] [CrossRef] [Green Version]
  59. He, Y.; Yan, L.; Zang, W.; Zheng, W. Novel sirtuin inhibitory warheads derived from the Nε-acetyl-lysine analog l-2-amino-7-carboxamidoheptanoic acid. Org. Biomol. Chem. 2015, 13, 10442–10450. [Google Scholar] [CrossRef]
  60. Polletta, L.; Vernucci, E.; Carnevale, I.; Arcangeli, T.; Rotili, D.; Palmerio, S.; Steegborn, C.; Nowak, T.; Schutkowski, M.; Pellegrini, L.; et al. SIRT5 regulation of ammonia-induced autophagy and mitophagy. Autophagy 2015, 11, 253–270. [Google Scholar] [CrossRef] [Green Version]
  61. Roessler, C.; Nowak, T.; Pannek, M.; Gertz, M.; Nguyen, G.T.T.; Scharfe, M.; Born, I.; Sippl, W.; Steegborn, C.; Schutkowski, M. Chemical Probing of the Human Sirtuin 5 Active Site Reveals Its Substrate Acyl Specificity and Peptide-Based Inhibitors. Angew. Chem. Int. Ed. 2014, 53, 10728–10732. [Google Scholar] [CrossRef]
  62. Kalbas, D.; Liebscher, S.; Nowak, T.; Meleshin, M.; Pannek, M.; Popp, C.; Alhalabi, Z.; Bordusa, F.; Sippl, W.; Steegborn, C.; et al. Potent and Selective Inhibitors of Human Sirtuin 5. J. Med. Chem. 2018, 61, 2460–2471. [Google Scholar] [CrossRef]
  63. Liu, J.; Zheng, W. Cyclic peptide-based potent human SIRT6 inhibitors. Org. Biomol. Chem. 2016, 14, 5928–5935. [Google Scholar] [CrossRef]
  64. Huang, Y.; Liu, J.; Yan, L.; Zheng, W. Simple N ε-thioacetyl-lysine-containing cyclic peptides exhibiting highly potent sirtuin inhibition. Bioorganic Med. Chem. Lett. 2016, 26, 1612–1617. [Google Scholar] [CrossRef]
  65. Jiang, Y.; Zheng, W. Cyclic Tripeptide-based Potent and Selective Human SIRT5 Inhibitors. Med. Chem. 2020, 16, 358–367. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, S.; Ji, S.; Yu, Z.-J.; Wang, H.-L.; Cheng, X.; Li, W.-J.; Jing, L.; Yu, Y.; Chen, Q.; Yang, L.-L.; et al. Structure-based discovery of new selective small-molecule sirtuin 5 inhibitors. Chem. Biol. Drug Des. 2017, 91, 257–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Suenkel, B.; Fischer, F.; Steegborn, C. Inhibition of the human deacylase Sirtuin 5 by the indole GW5074. Bioorganic Med. Chem. Lett. 2013, 23, 143–146. [Google Scholar] [CrossRef]
  68. Schuetz, A.; Min, J.; Antoshenko, T.; Wang, C.-L.; Allali-Hassani, A.; Dong, A.; Loppnau, P.; Vedadi, M.; Bochkarev, A.; Sternglanz, R.; et al. Structural Basis of Inhibition of the Human NAD+-Dependent Deacetylase SIRT5 by Suramin. Structure 2007, 15, 377–389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Maurer, B.; Rumpf, T.; Scharfe, M.; Stolfa, D.A.; Schmitt, M.L.; He, W.; Verdin, E.; Sippl, W.; Jung, M. Inhibitors of the NAD+-Dependent Protein Desuccinylase and Demalonylase Sirt5. ACS Med. Chem. Lett. 2012, 3, 1050–1053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Lugrin, J.; Ciarlo, E.; Santos, A.; Grandmaison, G.; Dos Santos, I.; Le Roy, D.; Roger, T. The sirtuin inhibitor cambinol impairs MAPK signaling, inhibits inflammatory and innate immune responses and protects from septic shock. Biochim. Biophys. Acta Mol. Cell Res. 2013, 1833, 1498–1510. [Google Scholar] [CrossRef] [Green Version]
  71. Orecchia, A.; Scarponi, C.; Di Felice, F.; Cesarini, E.; Avitabile, S.; Mai, A.; Mauro, M.L.; Sirri, V.; Zambruno, G.; Albanesi, C.; et al. Sirtinol Treatment Reduces Inflammation in Human Dermal Microvascular Endothelial Cells. PLoS ONE 2011, 6, e24307. [Google Scholar] [CrossRef]
  72. Han, H.; Li, C.; Li, M.; Yang, L.; Zhao, S.; Wang, Z.; Liu, H.; Liu, D. Design, Synthesis, and Biological Evaluation of 8-Mercapto-3,7-Dihydro-1H-Purine-2,6-Diones as Potent Inhibitors of SIRT1, SIRT2, SIRT3, and SIRT5. Molecules 2020, 25, 2755. [Google Scholar] [CrossRef]
  73. Yang, L.-L.; He, Y.-Y.; Chen, Q.-L.; Qian, S.; Wang, Z.-Y. Design and Synthesis of New 9-Substituted Norharmane Derivatives as Potential Sirt5 Inhibitors. J. Heterocycl. Chem. 2016, 54, 1457–1466. [Google Scholar] [CrossRef]
  74. Yang, L.; Wang, Z.; Qian, S.; He, Y.; Chen, Q. 9-Substituted Pyrido [3,4-b]Indole Derivative, ITS preparing Method, and its Application as SIRT Inhibitor, UNIVXIHUA. China Patent Application No. CN105884767, 11 January 2016. [Google Scholar]
  75. Yang, L.-L.; Wang, H.-L.; Yan, Y.-H.; Liu, S.; Yu, Z.-J.; Huang, M.-Y.; Luo, Y.; Zheng, X.; Yu, Y.; Li, G.-B. Sensitive fluorogenic substrates for sirtuin deacylase inhibitor discovery. Eur. J. Med. Chem. 2020, 192, 112201. [Google Scholar] [CrossRef]
  76. Walter, M.; Chen, I.P.; Vallejo-Gracia, A.; Kim, I.-J.; Bielska, O.; Lam, V.L.; Hayashi, J.M.; Cruz, A.; Shah, S.; Gross, J.D.; et al. SIRT5 is a proviral factor that interacts with SARS-CoV-2 Nsp14 protein. bioRxiv 2022. [Google Scholar] [CrossRef]
  77. Yang, L.; Peltier, R.; Zhang, M.; Song, D.; Huang, H.; Chen, G.; Chen, Y.; Zhou, F.; Hao, Q.; Bian, L.; et al. Desuccinylation-Triggered Peptide Self-Assembly: Live Cell Imaging of SIRT5 Activity and Mitochondrial Activity Modulation. J. Am. Chem. Soc. 2020, 142, 18150–18159. [Google Scholar] [CrossRef]
  78. Carafa, V.; Altucci, L.; Nebbioso, A. Dual Tumor Suppressor and Tumor Promoter Action of Sirtuins in Determining Malignant Phenotype. Front. Pharmacol. 2019, 10, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Mechanism of the deacetylation reaction catalyzed by SIRTs.
Scheme 1. Mechanism of the deacetylation reaction catalyzed by SIRTs.
Molecules 27 04449 sch001
Figure 1. (A) Ribbon diagram of SIRT5 (PDB code: 3RIY), evidencing the two domains and the position of the zinc cofactor. α-Helices are colored blue, β-sheets are yellow, while undefined loops are gray. (B) Surface representation of SIRT5 (PDB code: 3RIY), highlighting the succinyl-lysine peptide substrate (Thr-Ala-Arg-SLL-Ser-Thr-Gly-Gly, red) and the NAD+ cofactor (green), bound in their respective pockets. (C) Magnified view of the two binding pockets, highlighting the ligand (red), NAD+ (green), and the surrounding amino acids (light yellow).
Figure 1. (A) Ribbon diagram of SIRT5 (PDB code: 3RIY), evidencing the two domains and the position of the zinc cofactor. α-Helices are colored blue, β-sheets are yellow, while undefined loops are gray. (B) Surface representation of SIRT5 (PDB code: 3RIY), highlighting the succinyl-lysine peptide substrate (Thr-Ala-Arg-SLL-Ser-Thr-Gly-Gly, red) and the NAD+ cofactor (green), bound in their respective pockets. (C) Magnified view of the two binding pockets, highlighting the ligand (red), NAD+ (green), and the surrounding amino acids (light yellow).
Molecules 27 04449 g001
Figure 2. Graphical representation of the main metabolic pathways regulated by SIRT5. The activated targets are contained in green boxes, whereas inhibited proteins are in red boxes.
Figure 2. Graphical representation of the main metabolic pathways regulated by SIRT5. The activated targets are contained in green boxes, whereas inhibited proteins are in red boxes.
Molecules 27 04449 g002
Figure 3. PRISMA flow diagram, describing the screening process that led to the selection of the works discussed in this review.
Figure 3. PRISMA flow diagram, describing the screening process that led to the selection of the works discussed in this review.
Molecules 27 04449 g003
Figure 4. Chemical structures of the natural activators (17) of SIRT5.
Figure 4. Chemical structures of the natural activators (17) of SIRT5.
Molecules 27 04449 g004
Figure 5. Chemical structures of the synthetic activators (916) of SIRT5.
Figure 5. Chemical structures of the synthetic activators (916) of SIRT5.
Molecules 27 04449 g005
Figure 6. Chemical structures of the natural inhibitors (1723) of SIRT5.
Figure 6. Chemical structures of the natural inhibitors (1723) of SIRT5.
Molecules 27 04449 g006
Figure 7. Chemical structures of the linear peptides (2431) with inhibitory activity against SIRT5.
Figure 7. Chemical structures of the linear peptides (2431) with inhibitory activity against SIRT5.
Molecules 27 04449 g007
Figure 8. Chemical structures of the linear peptides (3239) with inhibitory activity against SIRT5.
Figure 8. Chemical structures of the linear peptides (3239) with inhibitory activity against SIRT5.
Molecules 27 04449 g008
Figure 9. Chemical structures of the cyclic peptides (4042) with inhibitory activity against SIRT5.
Figure 9. Chemical structures of the cyclic peptides (4042) with inhibitory activity against SIRT5.
Molecules 27 04449 g009
Figure 10. Chemical structures of the synthetic derivatives (4350) with inhibitory activity against SIRT5.
Figure 10. Chemical structures of the synthetic derivatives (4350) with inhibitory activity against SIRT5.
Molecules 27 04449 g010
Figure 11. Chemical structures of the synthetic derivatives (5159) with inhibitory activity against SIRT5.
Figure 11. Chemical structures of the synthetic derivatives (5159) with inhibitory activity against SIRT5.
Molecules 27 04449 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mori, M.; Cazzaniga, G.; Meneghetti, F.; Villa, S.; Gelain, A. Insights on the Modulation of SIRT5 Activity: A Challenging Balance. Molecules 2022, 27, 4449. https://doi.org/10.3390/molecules27144449

AMA Style

Mori M, Cazzaniga G, Meneghetti F, Villa S, Gelain A. Insights on the Modulation of SIRT5 Activity: A Challenging Balance. Molecules. 2022; 27(14):4449. https://doi.org/10.3390/molecules27144449

Chicago/Turabian Style

Mori, Matteo, Giulia Cazzaniga, Fiorella Meneghetti, Stefania Villa, and Arianna Gelain. 2022. "Insights on the Modulation of SIRT5 Activity: A Challenging Balance" Molecules 27, no. 14: 4449. https://doi.org/10.3390/molecules27144449

Article Metrics

Back to TopTop