Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 May 1.
Published in final edited form as: Cancer Res. 2016 Apr 20;76(9):2500–2506. doi: 10.1158/0008-5472.CAN-15-2733

Mitochondrial sirtuins in cancer: Emerging roles and therapeutic potential

Jasmine George 1, Nihal Ahmad 1,2
PMCID: PMC4874653  NIHMSID: NIHMS778394  PMID: 27197261

Abstract

The past few decades have witnessed a furious attention of scientific community towards identifying novel molecular factors and targets which could be exploited for drug development for cancer management. One such factor is the sirtuin (SIRT) family of nicotinamide adenine dinucleotide (NAD+)-dependent deacetylases. The role of SIRTs in cancer is extremely complex, with dichotomous functions depending on cell contexts. Mammalian SIRTs (SIRT1–7), differ in their cellular localization and biological functions. Amongst these, SIRT -3, -4, and -5 are located in the mitochondria and are being carefully investigated. These mitochondrial SIRTs (mtSIRTs) regulate multiple cellular and physiological processes, including cell cycle, gene expression, cell viability, stress response, metabolism and energy homeostasis. Recent research suggests that mtSIRTs influence tumors by regulating the metabolic state of the cell. While the research on the role of mtSIRTs in cancer is still in its infancy, studies have suggested tumor suppressor as well as tumor promoter roles for them. This review is focused on discussing an up-to-date information about the roles and functional relevance of mtSIRTs (SIRT -3, -4, -5) in cancers. We have also provided a critical discussion and our perspective on their dual roles, as tumor promoter versus tumor suppressor, in cancer.

Introduction

Cancer is a cluster of diseases with an underlying common feature of the development of abnormal cells that propagate out-of-control to travel and target almost every organ of the body to make it dysfunctional. According to American Cancer Society, it is expected that, the global cancer burden will nearly double by 2030, with 21.4 million cases and 13.2 million deaths (1). The last few decades have seen tremendous progress towards identifying the genes and pathways in cancer development and progression. The members of the Sir2 (silent information regulator 2) family, or sirtuins, have been widely investigates for their role in a variety of biological processes and disease conditions.

The sirtuins (SIRTs) belong to a family of class III histone deacetylases (HDACs) enzymes, which are highly conserved enzyme homologues of the yeast Sir2 protein, with nicotine adenine dinucleotide NAD+-dependent protein deacetylases and/or mono ADP ribosyltransferase activities (2). They are also known for regulating post-translational acyl modifications in organisms ranging from bacteria to human (3,4). Mammalian SIRTs make up a superfamily of seven SIRTs (SIRT1-SIRT7). Among these, SIRT1 is predominantly nuclear and SIRT2 is localized mainly in cytoplasm. However, they have the ability to shuttle between the nucleus and cytoplasm. SIRT3, SIRT4 and SIRT5 are primarily mitochondrial proteins, whereas SIRT6 and SIRT7 are nuclear sirtuins (5).

The SIRT family of proteins are becoming increasingly interesting and important due to the seemingly dichotomous role of this enzyme family in cancer biology, with both tumor promoter and tumor suppressor functions suggested in different cancers. In this review, we have focused on the role of mitochondrial sirtuins (mtSIRTs) in cancer. Indeed, mitochondrial cell signaling is extremely critical for cancer cells to grow and proliferate; and metabolic imbalances and enhanced resistance to mitochondrial apoptosis are well established hallmarks of cancer cells.

Discovery of sirtuin family of enzymes: an important class of histone deacetylases (HDACs)

The founding member of sirtuin family, Sir2, was first discovered as regulators of aging in the budding yeast Saccharomyces cerevisiae. In the late 1990s a study led by Matt Kaeberlein, verified that Sir2 overexpression prolongs yeast lifespan and deletion of Sir2 reduces yeast lifespan (6). However, a study by Imai and Guarente showed, for the first time, the true enzymatic activity of Sir2, as a NAD+ dependent histone deacetylase and gave a possible description of the mechanism by which Sir2 controls yeast aging (7). The evolutionarily conserved orthologues of Sir2 are shown to be found in most forms of life, including eukaryotes and prokaryotes (8). Structurally, the sirtuins members (SIRT1–SIRT7) share a conserved catalytic core of ~275 amino acids that is flanked by N- and C- terminal sequences of variable length. A small zinc binding domain and a large Rossmann-fold domain endows all SIRTs with the specificity to bind NAD+ as a cofactor. SIRTs enzymatic activities have been linked to a variety of cellular processes such as stress response, differentiation, metabolism, apoptosis, and cell survival due to their ability to deacetylate both histone and numerous non-histone targets (9). Although SIRTs have been studied for over a decade, the scientific community is still struggling to establish their complex role in cancer.

Mitochondrial SIRTs: the master regulators of metabolic homeostasis

As aforementioned, there are three known mtSIRTs (SIRT -3, -4, -5) shown to regulate a variety of cellular processes. SIRT3, a 45kDa protein, is located on chromosome 11p15.5 having an N-terminal mitochondrial targeting sequence that is cleaved off after import into the mitochondria, leaving an enzymatically active 28kDa protein. SIRT4 is 35.2kDa protein located on chromosome 12q24.31 whereas SIRT5 (33.9kDa)-like pseudogene is present on chromosome 1p31.2–32.1(8). Most of their functions are recently reviewed by Shih et al. (10). SIRT3 has a strong deacetylase activity whereas, SIRT4 has ADP ribosyltransferase activity, lipoamidase and biotinidase activities in mitochondria (11). SIRT5 has protein lysine desuccinylase, demalonylase, deglutarylase and a weak deacetylase activities. In addition, SIRT5 deacetylates carbamoyl phosphate synthetase 1 (CPS1), an enzyme catalyzing the initial step of the urea cycle for ammonia detoxification and disposal (12). All these three SIRTs are metabolically active in tissues like kidney, liver, adipose, heart, and brain.

The important role of mtSIRTs in cancer is not surprising owing to the fact that the process of cancer development and progression involves major alterations in cellular metabolism and mitochondrial energetics. Moreover, cancer cells are metabolically active and consume more cellular fuel than normal cells (13). Otto Warburg first proposed the theory in 1920s that defects in energy metabolism, particularly in mitochondrial function, may be the root cause of cancer. The proposed ‘Warburg effect’ suggests that cancer cells prefer glycolytic breakdown of glucose for energy, rather than mitochondrial oxidative phosphorylation (14). Mitochondria is considered as the energy provider and regulates various homeostatic processes, such as cell proliferation, apoptosis, oxidative stress, and calcium homeostasis. Understanding the mechanisms that link cellular energy metabolism and regulation of gene expression is of fundamental importance that could potentially be exploited for the management of various diseases, including cancer. The mtSIRTs are critical for the maintenance of cellular and mitochondria homeostasis through regulating mitochondria metabolism and protecting cells against oxidative stress via regulating reactive oxygen species (ROS) generation, a critical mechanism in cancer. Utilizing the mitochondrial NAD pool, the mtSIRTs can deacetylate targets implicated in the regulation of both glycolysis and cellular oxidative stress.

A more detailed discussion of mtSIRTs, including SIRT5 is elegantly provided in a very recent article by Kumar and Lombard (15). A brief account of mtSIRTs and their biological functions, especially relevant to cancer pathophysiology, is provided below.

SIRT3

SIRT3 is most well characterized and studied mtSIRT that has been shown to regulate almost every major aspect of mitochondrial biology (1618). Not surprisingly, SIRT3 is reported to contribute to the age-related diseases, metabolic, cardiovascular, neurogenerative diseases and cancer (19). SIRT3 deacetylates key mitochondrial proteins, including acetyl-coenzyme A synthetase 2, isocitrate dehydrogenase 2 (IDH2) and glutamate dehydrogenase (GDH) (20,21). Recently, SIRT3 is found to deacetylate and increase pyruvate dehydrogenase activity in cancer cells, which can increase both mitochondrial bioenergetics and glycolysis (22). Further, SIRT3 binds to and deacetylates mitochondrial pyruvate carrier 1 (MPC1) to promote its function, thereby inhibiting colon cancer cell growth (23). SIRT3 functions as a downstream target of peroxisome proliferator-activated receptor-gamma coactivator 1 alpha (PGC-1α) and decreases cellular ROS production and stimulate mitochondrial biogenesis (17). Indeed, PGC-1α, a transcriptional co-activator of genes involved in energy homeostasis, including peroxisome proliferator-activated receptor α (PPARs), estrogen-related receptor (ERRs) and nuclear respiratory factors (Nrfs), is a master regulator of metabolism and mitochondrial biogenesis.

A key metabolic enzyme that is directly regulated by SIRT3 in mitochondria is manganese superoxide (MnSOD) that protects the cells by detoxifying ROS (24). SIRT3 promotes antioxidant activity of MnSOD via direct deacetylation, and loss of SIRT3 increases acetylation of MnSOD to enhance ROS, which stabilizes hypoxia-inducible factor 1-alpha (HIF1-α), thereby facilitating tumor development (25). Conversely, Cui et al. found that overexpression of SIRT3 increases lactate production, ATP production leading to increased glycolysis which together with increased MnSOD activity, decreased ROS level to promoted cell proliferation of gastric cancer cells (26). In glioma cells, SIRT3 maintains the mitochondrial DNA integrity through its ability to deacetylate 8-oxoguanine-DNA glycosylase 1 (OGG1), thus preventing apoptosis and promoting carcinogenesis (27). Also, SIRT3 has been recently shown to deacetylates glutamate oxaloacetate transaminase (GOT2) that regulates the malate-aspartate NADH shuttle to control the growth of pancreatic tumor (28). Mitochondrial impairment and ROS production has been implicated in the development of insulin resistance and linked with diabetes. Consistent with this framework, decreased level of SIRT3 along with its associated transcription factor, forkhead box O3 (FOXO3a) was observed in mouse3T3L1 adipocytes and C2C12 myotubes upon chronic arsenic exposure (29). This decrease in SIRT3 caused enhanced ROS generation and mitochondrial membrane damage, leading to insulin resistance (29). Another study shows SIRT3 deficient mice when fed a chronic high-fat diet showed accelerated obesity, insulin resistance, and steatohepatitis, compared to wild-type mice (30).

SIRT4

SIRT4 has also been shown to regulate cellular metabolism by a variety of mechanisms. Glutamine metabolism is important for mitochondrial functions and macromolecular synthesis and could play an important role in cancer progression. SIRT4 has been shown to act as a tumor suppressor by regulating glutamine metabolism, suggesting a potential therapeutic use in glutamine-dependent neoplasms, such as B cell lymphoma (31). In addition, C-terminal-binding protein (CtBP)-mediated suppression of SIRT4 and enzymatic modification of GDH has been shown to promote glutaminolysis, in breast cancer cells (32). Also, mammalian target of rapamycin complex 1 (mTORC1) promotes mitochondrial glutamine metabolism by stimulating GDH and repressing the transcription of SIRT4 regulator cAMP-responsive element binding 2 (CREB2), and SIRT4 overexpression reduces proliferation and transformation of tuberous sclerosis 2 mutant cells and tumor development in a xenograft model (33). Besides, SIRT4 has also been implicated in other mitochondrial dysfunction related disorders like insulin secretion, obesity and cardiovascular diseases (3436). Therefore, based on published studies, SIRT4 appears to have a tumor suppressor function.

SIRT5

Phylogenetically most closely related to prokaryotic sirtuins, SIRT5 is predominantly mitochondrial, with some evidence of existence of functional extra-mitochondrial protein. Limited number of studies have examined the role of SIRT5 in cellular metabolic functions. SIRT5-mediated lysine desuccinylation has been shown to modulate a number of diverse metabolic pathways (37). Probably, the first clue suggesting the role of SIRT5 in metabolism and cancer comes from the observation that SIRT5 desuccinylates SOD1 and eliminates ROS thereby inhibits lung cancer cells growth (38). Additionally, SIRT5-mediated deacetylation of FOXO3 is shown to play an essential role in protecting lung epithelial cells from cigarette smoke extract-induced apoptosis (39). As described in a very recent review, SIRT5 deacetylates, desuccinylates, demalonylates, and/or deglutarylates multiple metabolic enzymes to activate or inhibit a variety of metabolic/cellular processes, including glucose oxidation, ketone body formation, fatty acid oxidation, ammonia detoxification, and ROS management (15).

Mitochondrial SIRTs and their molecular targets: Relevance to cancer

The research on mtSIRT’s role in cancer and its downstream targets has gained a lot of momentum in the recent years. A wide variety of molecular targets are associated with the different biological functions modulated by SIRTs. Some of these important targets are described below.

SIRT3

The fact that SIRT3 can regulate many hallmarks of the cancer cells, implicates SIRT3 as a potential therapeutic target for cancer management (40). However, the role of SIRT3 in cancer is being extensively debated, with numerous evidences supporting both tumor promoter and suppressor roles for SIRT3. SIRT3 was initially suggested to act as a tumor suppressor when introduction of a single oncogene (Myc or Ras) into the mouse embryonic fibroblasts (MEFs) from SIRT3-knockout mice resulted in an in vitro transformation (41). Further, Yang et al. reported that SIRT3 was aberrantly downregulated in gastric cancer tissue samples from patients (42). Recently, decreased SIRT3 expression was found in mantle cell lymphoma tissues and B cell malignant cells. This SIRT3 loss was found to correlate with hyperacetylation of IDH2 and SOD2, thereby decreasing IDH2 and SOD2 activities in B cell malignant cells (43). Another study demonstrated that SIRT3 is decreased in gastric cancer cells and its exogenous overexpression suppressed the proliferation of gastric cancer cells by downregulating Notch-1 (44). SIRT3 deacetylates and destabilizes the proto-oncogene product S-phase kinase-associated protein 2 (Skp2), and inactivation of SIRT3 leads to Skp2 acetylation and increased-stability and cytoplasmic retention, causing enhanced cellular proliferation, migration, and tumorigenesis (45). Song et al. demonstrated that SIRT3 inhibits hepatocellular carcinoma growth via glycogen synthase kinase-3β (GSK-3β)/Bax-dependent apoptotic pathway (46). In addition, binding of SIRT3 with nicotinamide mononucleotide adenylyltransferase 2 (NMNAT2) was shown to regulate cell proliferation and apoptosis in non-small cell lung carcinoma (NSCLC) cell lines, implicating the interaction between SIRT3 and NMNAT2, energy metabolism associated with SIRT3 (47). More recently, depletion of SIRT3 was shown to exert tumor-suppressive function by regulating iron metabolism in pancreatic cancer (48). Also, exogenous overexpression of SIRT3 was shown to suppress prostate cancer by decreasing ROS generation which inhibits Akt phosphorylation at Ser 473 leading to ubiquitin-mediated proteosomal c-MYC degradation (49).

In certain cancers, SIRT3 is found to be overexpressed (5052) and associated with therapy resistance (53). Interestingly, both tumor suppressor and oncogenic function of SIRT3 have been reported in breast cancer (54,55). Liu et al. have shown a transcriptional activation of SIRT3 in several radiation-treated human tumor cells and their corresponding xenograft tumors that was mediated via NF-κB and Thr150/Ser159 phosphorylation by cyclin B1-CDK1 (56). In another study, clinically, high cytoplasmic SIRT3 expression was shown to correlate with high tumor grades, positive lymph node status, and poor prognosis in colon cancer patients (57). Among 127 patients, high SIRT3 expression was seen in 71 patients and low SIRT3 expression was seen in 56 patients. The overall survival rate was 80.2% among patients with low SIRT3 expressions and 55.9% among patients with high SIRT3 expressions (log-rank P = 0.002). However, the relevance of cytoplasmic SIRT3 needs to be carefully investigated in future studies. In a very recent study from our laboratory, we found that SIRT3 is overexpressed in human melanoma cells and clinical melanoma tissues (58). Further, a short hairpin RNA (shRNA)-mediated knockdown of SIRT3 was found to impart significant anti-proliferative effects in multiple human melanoma cell lines (58). Conversely, forced exogenous overexpression of SIRT3 promoted enhanced proliferative potential of Hs294T melanoma cells and normal immortalized Mel-ST melanocytes (58). In addition, SIRT3 knockdown also resulted in a decrease in tumor growth in vivo in a xenograft model (58). In addition, limited information is available regarding mutation(s) in SIRT3 and its role in cancer. Chen et al. have found that point mutation in SIRT3 reduced its deacetylase activity promoting proliferation of oral squamous cell carcinoma (OSCC) cells (59). However, the exact role of mutant SIRT3 in tumorigenesis needs to be further explored.

SIRT3 has been shown to target and deacetylate the lysine residues of not only mitochondrial proteins but numerous other proteins as well. An interaction between SIRT3 and tumor suppressor p53 has been shown in certain cancers. SIRT3 was shown to partially abrogate p53 activity to enact growth arrest and senescence in bladder carcinoma (60). Similarly, an exogenous overexpression of SIRT3 was found to inhibit HCC cell growth by reducing mouse double minute 2 (MDM2)-mediated p53 degradation, possibly via post-transcriptional regulation (61). Interestingly, SIRT3 overexpression upregulated p53 and its downstream factor p21 and decreased ROS level in lung adenocarcinoma cells thereby imparting pro-proliferative response (62). In cardiomyocytes, SIRT3 deacetylates Ku70, augmenting Ku70-Bax interactions and consequently preventing Bax translocation to the mitochondria, thus protecting these cells from genotoxic and oxidative stress-mediated cell death to promote cell survival (63). Thus, it appears that SIRT3 can have oncogenic or tumor suppressive functions, based on cancer types and both known and possibly yet unidentified downstream targets.

SIRT4

Limited information is available regarding the downstream targets of SIRT4 and its interactions with other pathways that affect metabolic reprogramming in cancer cells. As discussed above, a few reports articulate tumor-suppressive activities of SIRT4 via its effects on mitochondrial glutamine metabolism, in part through modification and repression of GDH. In a recent study, down-regulation of SIRT4 has been shown to be associated with poor prognosis in head and neck squamous cell carcinoma (HNSCC) patients (64). Tumor suppressor function of SIRT4 is also reported in colorectal cancer where SIRT4 has been shown to cause E-cadherin upregulation thereby suppressing the proliferation, migration and invasion through inhibition of glutamine metabolism (65). Jeong et al. have shown that SIRT4 loss causes increased glutamine-dependent proliferation and stress-induced genomic instability, and SIRT4 knockout mice spontaneously develop lung tumors (66). Huang et al. have found a downregulation of SIRT4 in gastric adenocarcinoma, suggesting its tumor suppressor function (67). Thus, SIRT4 loss or inactivation seems to create a tumor-permissive environment to provide metabolic advantage, supporting cell proliferation.

SIRT5

Like SIRT4, SIRT5 is also not extensively studied and characterized, especially in context with cancer. SIRT5 has been found to be overexpressed in human NSCLC and can facilitate drug resistance in in vitro and in vivo by reducing expression of Nrf2 and its downstream targets (68). Poletta et al. have shown SIRT5 regulation of ammonia-induced autophagy and mitophagy in breast cancer cells (69). Unlike in NSCLC, SIRT5 has been shown to be considerably downregulated in HNSCC tissues compared with noncancerous tissues, suggesting its possible tumor suppressor function (70). Guan et al. have identified SIRT5 as a deacetylases for the tumor suppressor promyelocytic leukemia protein (PML), suggesting a tumor suppressor function for this SIRT (71).

Thus, SIRT3 and SIRT5 have been shown to exhibit tumor suppressor as well as tumor promoter properties under different cellular conditions, tumor stage and tissue of origin. Interestingly, based on limited number of studies, SIRT4 appears to have only a tumor suppressor function. However, further detailed studies are needed to dissect the downstream targets of mtSIRTs.

Activators and inhibitors of mitochondrial SIRTs: Potential uses in cancer therapy

Recent research has established the essential roles of mtSIRTs in a wide range of critical mitochondrial functions and cellular processes. This has spurred extensive research and efforts from scientists across multiple disciplines towards the development of small-molecule modulators (activators and inhibitors) of these SIRTs. Although considerable efforts have been made to develop sirtuin inhibitors and activators (Reviewed in (72,73)) in the recent past, only limited success has been achieved. Majority of the know modulators of sirtuins seem to suffer from specificity issues. To our knowledge, no specific modulators of mtSIRTs have been reported. Thus, this area of research is still in infancy and needs a concerted attention from scientists and drug companies. Table 1 provides a list of recently identified and characterized mtSIRT-modulators together with their IC50 values.

Table 1.

Putative modulators (activators and inhibitors) for mitochondrial sirtuins*.

Sirtuin Putative modulators IC50 value References
SIRT3 Resveratrol (activator and inhibitor) - (75,76)
Adjudin (activator) - (77)
Oroxylin A (activator) - (78)
Honokiol (activator) - (79)
Pyrroloquinolinequinone (activator) - (80)
LC-0296 (inhibitor) 3.6 µmol/L (81)
EX-527 (inhibitor) 48.7 µmol/L (82)
AGK2 (inhibitor) >50 µmol/L (83)
Nicotinamide (inhibitor) 36.7±1.3 µmol/L (84)
SRT1720 (inhibitor) 11±1 mmol/L (85)
3-(1H-1,2,3-triazol-4-yl)pyridine (inhibitor) 38 ± 5 µmol/L (86)
4-Bromo’-resveratrol (inhibitor) 143± 3.6 µmol/L (76)
Thieno[3,2-d]pyrimidine-6-carboxamide (inhibitor) 33 nmol/L (87)
SDX-437 (inhibitor) 700 nmol/L (88)
Tenovin-6 (inhibitor) 67 µmol/L (89)
SIRT4 No report -
SIRT5 Resveratrol (activator) - (90)
Suramin (inhibitor) 22 µmol/L (91)
Nε-carboxyethyl-thiocarbamoyl-lysine (inhibitor) 5.0 ± 1.9 µmol/L (92)
GW5074 (inhibitor) 19.5 ± 7.3 µmol/L (93)
*

This table provides a list of most of the known activators/inhibitors of mitochondrial sirtuins and their reported IC50 values. The IC50 values represent deacetylase activities identified in biochemical assays, inhibiting ~50% dacetylation of synthetic peptides. This table represents some better studied examples and may not be a complete list of activators/inhibitors of mitochondrial sirtuins.

Concluding discussion and perspective

Mitochondrial SIRTs are critical regulators of metabolic functions in the cells, which are responsible for keeping the cellular homeostasis and cellular energetics ‘in check’. Emerging evidence are now suggesting that mtSIRTs are important regulator of what is known as “oncometabolism”, defined as the ensemble of metabolic rearrangements that accompany oncogenesis and tumor progression (74). Increasing understanding of the function of mtSIRTs is providing an insights on their functional significance in cancer. The mtSIRTs have been shown to target a variety of metabolic substrates that are relevant to pathophysiology of cancer. A description of the key metabolic substrates of mtSIRTs, their biological functions, downstream targets and expression pattern in cancer is provided in Table 2. It seems possible that the SIRTs may be useful as novel biomarkers for cancer diagnoses and targets for development of novel strategies for cancer therapy. However, further research is needed to answer the question, why and how these sirtuins have both oncogenic or tumor-suppressive role, in different cancer and how this dual action could be best exploited in cancer management? It seems probable that the role of sirtuins is different in cancer cells versus normal cells and/or depends on the stage of neoplastic transformation. Future research needs to address these possibilities. As discussed above, there is a need to intensify our efforts towards the development of novel specific pharmacological activators and inhibitors of SIRTs that can serve as a tool to understand the sirtuin biology and may also possibly be useful in the management of diseases where SIRTs are dysregulated, including cancer. However, given the important role of mtSIRTs in metabolism, there is always an issue of off-target complications. This needs to be carefully investigated. Further, there may be some cancers where more than one sirtuins are overexpressed (or downregulated). In these cases, the non-specific broad sirtuin inhibitors may be more useful, rather than specific ones. On the same line, additional research is required to ascertain the functional redundancy and overlaps among different sirtuins. This is particularly important for the mtSIRTs. In conclusion, the mtSIRTs have emerged as critical regulators of diverse metabolic functions relevant to cancer. Therefore, manipulating these enzymes may have important implications in cancer management.

Table 2.

Mitochondrial sirtuins: Metabolic substrates, functions, downstream targets, and expression pattern in various cancers*

Sirtuin Activity Substrates Functions Downstream
targets
Relative
expression in
various cancer
SIRT3
  • Deacetylates

AceCS2,
IDH2,
GDH,
MCP1,
PDH,
MnSOD,
OGG1,
GOT2,
FOXO3a,
Skp2
Metabolism/gene
transcription/
apoptosis/ DNA
repair/ oxidative
stress
PGC-1α,
Notch-1,
HIF1-α,
GSK-
3β/Bax,
NMNAT2,
PI3K/Akt,
NF-κB, p53,
p21, Ku70
  • Downregulated in gastric cancer, B cell malignancies, HCC, NSCLC, prostate cancer, breast cancer, pancreatic cancer

  • Upregulated in breast cancer, melanoma, colon cancer, OSCC, lung adenocarcinoma

SIRT4
  • ADP ribosyltransferase

  • Lipoamidase

  • Biotinidase

GDH Metabolism E-cadherin
  • Downregulated in B cell lymphoma, breast cancer, HNSCC, colorectal cancer, lung tumors in mice, gastric adenocarcinoma

SIRT5
  • Lysine desuccinylase,

  • Demalonylase,

  • Deglutarylase,

  • Deacetylates

CPS1,
PML
Metabolism Nrf2
  • Downregulated in HNSCC, leukemia

  • Upregulated in NSCLC

*

This table provides examples that have been studied in a greater detail. It is by no means a complete depiction of all the substrates, functions, downstream targets, and expression pattern in cancer.

Acknowledgments

This work was partially supported by funding from the NIH (R01AR059130 and R01CA176748) and the Department of Veterans Affairs (VA Merit Review Award 1I01BX001008).

Footnotes

Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

References

  • 1.Siegel RL, Miller KD, Jemal A. Cancer statistics, 2015. CA Cancer J Clin. 2015;65(1):5–29. doi: 10.3322/caac.21254. [DOI] [PubMed] [Google Scholar]
  • 2.North BJ, Verdin E. Sirtuins: Sir2-related NAD-dependent protein deacetylases. Genome Biol. 2004;5(5):224. doi: 10.1186/gb-2004-5-5-224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Thao S, Escalante-Semerena JC. Control of protein function by reversible Nvarepsilon-lysine acetylation in bacteria. Curr Opin Microbiol. 2011;14(2):200–204. doi: 10.1016/j.mib.2010.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Chen Y, Sprung R, Tang Y, Ball H, Sangras B, Kim SC, et al. Lysine propionylation and butyrylation are novel post-translational modifications in histones. Mol Cell Proteomics. 2007;6(5):812–819. doi: 10.1074/mcp.M700021-MCP200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Vassilopoulos A, Fritz KS, Petersen DR, Gius D. The human sirtuin family: evolutionary divergences and functions. Hum Genomics. 2011;5(5):485–496. doi: 10.1186/1479-7364-5-5-485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Kaeberlein M, McVey M, Guarente L. The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms. Genes Dev. 1999;13(19):2570–2580. doi: 10.1101/gad.13.19.2570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Imai S, Armstrong CM, Kaeberlein M, Guarente L. Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature. 2000;403(6771):795–800. doi: 10.1038/35001622. [DOI] [PubMed] [Google Scholar]
  • 8.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. Biochem Biophys Res Commun. 1999;260(1):273–279. doi: 10.1006/bbrc.1999.0897. [DOI] [PubMed] [Google Scholar]
  • 9.Michan S, Sinclair D. Sirtuins in mammals: insights into their biological function. Biochem J. 2007;404(1):1–13. doi: 10.1042/BJ20070140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Shih J, Donmez G. Mitochondrial sirtuins as therapeutic targets for age-related disorders. Genes Cancer. 2013;4(3–4):91–96. doi: 10.1177/1947601912474931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Mathias RA, Greco TM, Oberstein A, Budayeva HG, Chakrabarti R, Rowland EA, et al. Sirtuin 4 is a lipoamidase regulating pyruvate dehydrogenase complex activity. Cell. 2014;159(7):1615–1625. doi: 10.1016/j.cell.2014.11.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Yang X, Liu B, Zhu W, Luo J. SIRT5, functions in cellular metabolism with a multiple enzymatic activities. Sci China Life Sci. 2015;58(9):912–914. doi: 10.1007/s11427-015-4902-8. [DOI] [PubMed] [Google Scholar]
  • 13.Cantor JR, Sabatini DM. Cancer cell metabolism: one hallmark, many faces. Cancer Discov. 2012;2(10):881–898. doi: 10.1158/2159-8290.CD-12-0345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Warburg O. On the origin of cancer cells. Science. 1956;123(3191):309–314. doi: 10.1126/science.123.3191.309. [DOI] [PubMed] [Google Scholar]
  • 15.Kumar S, Lombard DB. Mitochondrial sirtuins and their relationships with metabolic disease and cancer. Antioxid Redox Signal. 2015;22(12):1060–1077. doi: 10.1089/ars.2014.6213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.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. Proc Natl Acad Sci U S A. 2008;105(38):14447–14452. doi: 10.1073/pnas.0803790105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Kong X, Wang R, Xue Y, Liu X, Zhang H, Chen Y, et al. Sirtuin 3, a new target of PGC-1alpha, plays an important role in the suppression of ROS and mitochondrial biogenesis. PLoS One. 2010;5(7):e11707. doi: 10.1371/journal.pone.0011707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.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. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.McDonnell E, Peterson BS, Bomze HM, Hirschey MD. SIRT3 regulates progression and development of diseases of aging. Trends Endocrinol Metab. 2015;26(9):486–492. doi: 10.1016/j.tem.2015.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Lombard DB, Alt FW, Cheng HL, Bunkenborg J, Streeper RS, Mostoslavsky R, et al. Mammalian Sir2 homolog SIRT3 regulates global mitochondrial lysine acetylation. Mol Cell Biol. 2007;27(24):8807–8814. doi: 10.1128/MCB.01636-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Papa L, Germain D. SirT3 regulates the mitochondrial unfolded protein response. Mol Cell Biol. 2014;34(4):699–710. doi: 10.1128/MCB.01337-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Ozden O, Park SH, Wagner BA, Yong Song H, Zhu Y, Vassilopoulos A, et al. SIRT3 deacetylates and increases pyruvate dehydrogenase activity in cancer cells. Free Radic Biol Med. 2014;76:163–172. doi: 10.1016/j.freeradbiomed.2014.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Liang L, Li Q, Huang L, Li D, Li X. Sirt3 binds to and deacetylates mitochondrial pyruvate carrier 1 to enhance its activity. Biochem Biophys Res Commun. 2015;468(4):807–812. doi: 10.1016/j.bbrc.2015.11.036. [DOI] [PubMed] [Google Scholar]
  • 24.Tao R, Coleman MC, Pennington JD, Ozden O, Park SH, Jiang H, et al. Sirt3-mediated deacetylation of evolutionarily conserved lysine 122 regulates MnSOD activity in response to stress. Mol Cell. 2010;40(6):893–904. doi: 10.1016/j.molcel.2010.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Finley LW, Carracedo A, Lee J, Souza A, Egia A, Zhang J, et al. SIRT3 opposes reprogramming of cancer cell metabolism through HIF1alpha destabilization. Cancer Cell. 2011;19(3):416–428. doi: 10.1016/j.ccr.2011.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Cui Y, Qin L, Wu J, Qu X, Hou C, Sun W, et al. SIRT3 Enhances Glycolysis and Proliferation in SIRT3-Expressing Gastric Cancer Cells. PLoS One. 2015;10(6):e0129834. doi: 10.1371/journal.pone.0129834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Cheng Y, Ren X, Gowda AS, Shan Y, Zhang L, Yuan YS, et al. Interaction of Sirt3 with OGG1 contributes to repair of mitochondrial DNA and protects from apoptotic cell death under oxidative stress. Cell Death Dis. 2013;4:e731. doi: 10.1038/cddis.2013.254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Yang H, Zhou L, Shi Q, Zhao Y, Lin H, Zhang M, et al. SIRT3-dependent GOT2 acetylation status affects the malate-aspartate NADH shuttle activity and pancreatic tumor growth. EMBO J. 2015;34(8):1110–1125. doi: 10.15252/embj.201591041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Padmaja Divya S, Pratheeshkumar P, Son YO, Vinod Roy R, Andrew Hitron J, Kim D, et al. Arsenic Induces Insulin Resistance in Mouse Adipocytes and Myotubes Via Oxidative Stress-Regulated Mitochondrial Sirt3-FOXO3a Signaling Pathway. Toxicol Sci. 2015;146(2):290–300. doi: 10.1093/toxsci/kfv089. [DOI] [PubMed] [Google Scholar]
  • 30.Hirschey MD, Shimazu T, Jing E, Grueter CA, Collins AM, Aouizerat B, et al. SIRT3 deficiency and mitochondrial protein hyperacetylation accelerate the development of the metabolic syndrome. Mol Cell. 2011;44(2):177–190. doi: 10.1016/j.molcel.2011.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Jeong SM, Lee A, Lee J, Haigis MC. SIRT4 protein suppresses tumor formation in genetic models of Myc-induced B cell lymphoma. J Biol Chem. 2014;289(7):4135–4144. doi: 10.1074/jbc.M113.525949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Wang L, Zhou H, Wang Y, Cui G, Di LJ. CtBP maintains cancer cell growth and metabolic homeostasis via regulating SIRT4. Cell Death Dis. 2015;6:e1620. doi: 10.1038/cddis.2014.587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.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. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Liu B, Che W, Xue J, Zheng C, Tang K, Zhang J, et al. SIRT4 prevents hypoxia-induced apoptosis in H9c2 cardiomyoblast cells. Cell Physiol Biochem. 2013;32(3):655–662. doi: 10.1159/000354469. [DOI] [PubMed] [Google Scholar]
  • 35.Savastano S, Di Somma C, Colao A, Barrea L, Orio F, Finelli C, et al. Preliminary data on the relationship between circulating levels of Sirtuin 4, anthropometric and metabolic parameters in obese subjects according to growth hormone/insulin-like growth factor-1 status. Growth Horm IGF Res. 2015;25(1):28–33. doi: 10.1016/j.ghir.2014.10.006. [DOI] [PubMed] [Google Scholar]
  • 36.Ahuja N, Schwer B, Carobbio S, Waltregny D, North BJ, Castronovo V, et al. Regulation of insulin secretion by SIRT4, a mitochondrial ADP-ribosyltransferase. J Biol Chem. 2007;282(46):33583–33592. doi: 10.1074/jbc.M705488200. [DOI] [PubMed] [Google Scholar]
  • 37.Park J, Chen Y, Tishkoff DX, Peng C, Tan M, Dai L, et al. SIRT5-mediated lysine desuccinylation impacts diverse metabolic pathways. Mol Cell. 2013;50(6):919–930. doi: 10.1016/j.molcel.2013.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Lin ZF, Xu HB, Wang JY, Lin Q, Ruan Z, Liu FB, et al. SIRT5 desuccinylates and activates SOD1 to eliminate ROS. Biochem Biophys Res Commun. 2013;441(1):191–195. doi: 10.1016/j.bbrc.2013.10.033. [DOI] [PubMed] [Google Scholar]
  • 39.Wang Y, Zhu Y, Xing S, Ma P, Lin D. SIRT5 prevents cigarette smoke extract-induced apoptosis in lung epithelial cells via deacetylation of FOXO3. Cell Stress Chaperones. 2015;20(5):805–810. doi: 10.1007/s12192-015-0599-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Alhazzazi TY, Kamarajan P, Verdin E, Kapila YL. Sirtuin-3 (SIRT3) and the Hallmarks of Cancer. Genes Cancer. 2013;4(3–4):164–171. doi: 10.1177/1947601913486351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Kim HS, Patel K, Muldoon-Jacobs K, Bisht KS, Aykin-Burns N, Pennington JD, et al. SIRT3 is a mitochondria-localized tumor suppressor required for maintenance of mitochondrial integrity and metabolism during stress. Cancer Cell. 2010;17(1):41–52. doi: 10.1016/j.ccr.2009.11.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Yang B, Fu X, Shao L, Ding Y, Zeng D. Aberrant expression of SIRT3 is conversely correlated with the progression and prognosis of human gastric cancer. Biochem Biophys Res Commun. 2014;443(1):156–160. doi: 10.1016/j.bbrc.2013.11.068. [DOI] [PubMed] [Google Scholar]
  • 43.Yu W, Denu RA, Kruatkramer KA, Grindle KM, Yang DT, Asimakopoulos F, et al. Loss of SIRT3 Provides Growth Advantage for B Cell Malignancies. J Biol Chem. 2015;291(7):3268–3279. doi: 10.1074/jbc.M115.702076. [PubMed: 26631723] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Wang L, Wang WY, Cao LP. SIRT3 inhibits cell proliferation in human gastric cancer through down-regulation of Notch-1. Int J Clin Exp Med. 2015;8(4):5263–5271. [PMC free article] [PubMed] [Google Scholar]
  • 45.Inuzuka H, Gao D, Finley LW, Yang W, Wan L, Fukushima H, et al. Acetylation-dependent regulation of Skp2 function. Cell. 2012;150(1):179–193. doi: 10.1016/j.cell.2012.05.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Song CL, Tang H, Ran LK, Ko BC, Zhang ZZ, Chen X, et al. Sirtuin 3 inhibits hepatocellular carcinoma growth through the glycogen synthase kinase-3beta/BCL2-associated X protein-dependent apoptotic pathway. Oncogene. 2016;35(5):631–641. doi: 10.1038/onc.2015.121. [DOI] [PubMed] [Google Scholar]
  • 47.Li H, Feng Z, Wu W, Li J, Zhang J, Xia T. SIRT3 regulates cell proliferation and apoptosis related to energy metabolism in non-small cell lung cancer cells through deacetylation of NMNAT2. Int J Oncol. 2013;43(5):1420–1430. doi: 10.3892/ijo.2013.2103. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 48.Jeong SM, Lee J, Finley LW, Schmidt PJ, Fleming MD, Haigis MC. SIRT3 regulates cellular iron metabolism and cancer growth by repressing iron regulatory protein 1. Oncogene. 2015;34(16):2115–2124. doi: 10.1038/onc.2014.124. [PubMed: 24909164] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Quan Y, Wang N, Chen Q, Xu J, Cheng W, Di M, et al. SIRT3 inhibits prostate cancer by destabilizing oncoprotein c-MYC through regulation of the PI3K/Akt pathway. Oncotarget. 2015;6(28):26494–26507. doi: 10.18632/oncotarget.4764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Yan SM, Han X, Han PJ, Chen HM, Huang LY, Li Y. SIRT3 is a novel prognostic biomarker for esophageal squamous cell carcinoma. Med Oncol. 2014;31(8):103. doi: 10.1007/s12032-014-0103-8. [DOI] [PubMed] [Google Scholar]
  • 51.Alhazzazi TY, Kamarajan P, Joo N, Huang JY, Verdin E, D'Silva NJ, et al. Sirtuin-3 (SIRT3), a novel potential therapeutic target for oral cancer. Cancer. 2011;117(8):1670–1678. doi: 10.1002/cncr.25676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Shackelford R, Hirsh S, Henry K, Abdel-Mageed A, Kandil E, Coppola D. Nicotinamide phosphoribosyltransferase and SIRT3 expression are increased in well-differentiated thyroid carcinomas. Anticancer Res. 2013;33(8):3047–3052. [PubMed] [Google Scholar]
  • 53.Zhang L, Ren X, Cheng Y, Huber-Keener K, Liu X, Zhang Y, et al. Identification of Sirtuin 3, a mitochondrial protein deacetylase, as a new contributor to tamoxifen resistance in breast cancer cells. Biochem Pharmacol. 2013;86(6):726–733. doi: 10.1016/j.bcp.2013.06.032. [DOI] [PubMed] [Google Scholar]
  • 54.Ashraf N, Zino S, Macintyre A, Kingsmore D, Payne AP, George WD, et al. Altered sirtuin expression is associated with node-positive breast cancer. Br J Cancer. 2006;95(8):1056–1061. doi: 10.1038/sj.bjc.6603384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Desouki MM, Doubinskaia I, Gius D, Abdulkadir SA. Decreased mitochondrial SIRT3 expression is a potential molecular biomarker associated with poor outcome in breast cancer. Hum Pathol. 2014;45(5):1071–1077. doi: 10.1016/j.humpath.2014.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Liu R, Fan M, Candas D, Qin L, Zhang X, Eldridge A, et al. CDK1-Mediated SIRT3 Activation Enhances Mitochondrial Function and Tumor Radioresistance. Mol Cancer Ther. 2015;14(9):2090–2102. doi: 10.1158/1535-7163.MCT-15-0017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Liu C, Huang Z, Jiang H, Shi F. The sirtuin 3 expression profile is associated with pathological and clinical outcomes in colon cancer patients. Biomed Res Int. 2014;2014:871263. doi: 10.1155/2014/871263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.George J, Nihal M, Singh CK, Zhong W, Liu X, Ahmad N. Pro-proliferative function of mitochondrial sirtuin deacetylase SIRT3 in human melanoma. J Invest Dermatol. 2016;136(4):809–818. doi: 10.1016/j.jid.2015.12.026. [PubMed: 26743598] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Chen IC, Chiang WF, Liu SY, Chen PF, Chiang HC. Role of SIRT3 in the regulation of redox balance during oral carcinogenesis. Mol Cancer. 2013;12:68. doi: 10.1186/1476-4598-12-68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Li S, Banck M, Mujtaba S, Zhou MM, Sugrue MM, Walsh MJ. p53-induced growth arrest is regulated by the mitochondrial SirT3 deacetylase. PLoS One. 2010;5(5):e10486. doi: 10.1371/journal.pone.0010486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Zhang YY, Zhou LM. Sirt3 inhibits hepatocellular carcinoma cell growth through reducing Mdm2-mediated p53 degradation. Biochem Biophys Res Commun. 2012;423(1):26–31. doi: 10.1016/j.bbrc.2012.05.053. [DOI] [PubMed] [Google Scholar]
  • 62.Xiao K, Jiang J, Wang W, Cao S, Zhu L, Zeng H, et al. Sirt3 is a tumor suppressor in lung adenocarcinoma cells. Oncol Rep. 2013;30(3):1323–1328. doi: 10.3892/or.2013.2604. [DOI] [PubMed] [Google Scholar]
  • 63.Sundaresan NR, Samant SA, Pillai VB, Rajamohan SB, Gupta MP. SIRT3 is a stress-responsive deacetylase in cardiomyocytes that protects cells from stress-mediated cell death by deacetylation of Ku70. Mol Cell Biol. 2008;28(20):6384–6401. doi: 10.1128/MCB.00426-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Mahjabeen I, Kayani MA. Loss of Mitochondrial Tumor Suppressor Genes Expression Is Associated with Unfavorable Clinical Outcome in Head and Neck Squamous Cell Carcinoma: Data from Retrospective Study. PLoS One. 2016;11(1):e0146948. doi: 10.1371/journal.pone.0146948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Miyo M, Yamamoto H, Konno M, Colvin H, Nishida N, Koseki J, et al. Tumour-suppressive function of SIRT4 in human colorectal cancer. Br J Cancer. 2015;113(3):492–499. doi: 10.1038/bjc.2015.226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.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. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Huang G, Cui F, Yu F, Lu H, Zhang M, Tang H, et al. Sirtuin-4 (SIRT4) is downregulated and associated with some clinicopathological features in gastric adenocarcinoma. Biomed Pharmacother. 2015;72:135–139. doi: 10.1016/j.biopha.2015.04.013. [DOI] [PubMed] [Google Scholar]
  • 68.Lu W, Zuo Y, Feng Y, Zhang M. SIRT5 facilitates cancer cell growth and drug resistance in non-small cell lung cancer. Tumour Biol. 2014;35(11):10699–10705. doi: 10.1007/s13277-014-2372-4. [DOI] [PubMed] [Google Scholar]
  • 69.Polletta L, Vernucci E, Carnevale I, Arcangeli T, Rotili D, Palmerio S, et al. SIRT5 regulation of ammonia-induced autophagy and mitophagy. Autophagy. 2015;11(2):253–270. doi: 10.1080/15548627.2015.1009778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Lai CC, Lin PM, Lin SF, Hsu CH, Lin HC, Hu ML, et al. Altered expression of SIRT gene family in head and neck squamous cell carcinoma. Tumour Biol. 2013;34(3):1847–1854. doi: 10.1007/s13277-013-0726-y. [DOI] [PubMed] [Google Scholar]
  • 71.Guan D, Lim JH, Peng L, Liu Y, Lam M, Seto E, et al. Deacetylation of the tumor suppressor protein PML regulates hydrogen peroxide-induced cell death. Cell Death Dis. 2014;5:e1340. doi: 10.1038/cddis.2014.185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Chen L. Medicinal chemistry of sirtuin inhibitors. Curr Med Chem. 2011;18(13):1936–1946. doi: 10.2174/092986711795590057. [DOI] [PubMed] [Google Scholar]
  • 73.Villalba JM, Alcain FJ. Sirtuin activators and inhibitors. Biofactors. 2012;38(5):349–359. doi: 10.1002/biof.1032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Galluzzi L, Kroemer G. Preface: oncometabolism: a new field of research with profound therapeutic implications. Methods Enzymol. 2014;542:xix–xxiii. doi: 10.1016/B978-0-12-416618-9.10000-7. [DOI] [PubMed] [Google Scholar]
  • 75.Chen T, Li J, Liu J, Li N, Wang S, Liu H, et al. Activation of SIRT3 by resveratrol ameliorates cardiac fibrosis and improves cardiac function via the TGF-beta/Smad3 pathway. Am J Physiol Heart Circ Physiol. 2015;308(5):H424–H434. doi: 10.1152/ajpheart.00454.2014. [DOI] [PubMed] [Google Scholar]
  • 76.Nguyen GT, Gertz M, Steegborn C. Crystal structures of Sirt3 complexes with 4'-bromo-resveratrol reveal binding sites and inhibition mechanism. Chem Biol. 2013;20(11):1375–1385. doi: 10.1016/j.chembiol.2013.09.019. [DOI] [PubMed] [Google Scholar]
  • 77.Quan Y, Xia L, Shao J, Yin S, Cheng CY, Xia W, et al. Adjudin protects rodent cochlear hair cells against gentamicin ototoxicity via the SIRT3-ROS pathway. Sci Rep. 2015;5:8181. doi: 10.1038/srep08181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Wei L, Zhou Y, Dai Q, Qiao C, Zhao L, Hui H, et al. Oroxylin A induces dissociation of hexokinase II from the mitochondria and inhibits glycolysis by SIRT3-mediated deacetylation of cyclophilin D in breast carcinoma. Cell Death Dis. 2013;4:e601. doi: 10.1038/cddis.2013.131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Pillai VB, Samant S, Sundaresan NR, Raghuraman H, Kim G, Bonner MY, et al. Honokiol blocks and reverses cardiac hypertrophy in mice by activating mitochondrial Sirt3. Nat Commun. 2015;6:6656. doi: 10.1038/ncomms7656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Zhang J, Meruvu S, Bedi YS, Chau J, Arguelles A, Rucker R, et al. Pyrroloquinoline quinone increases the expression and activity of Sirt1 and -3 genes in HepG2 cells. Nutr Res. 2015;35(9):844–849. doi: 10.1016/j.nutres.2015.06.014. [DOI] [PubMed] [Google Scholar]
  • 81.Alhazzazi TY, Kamarajan P, Xu Y, Ai T, Chen L, Verdin E, et al. A Novel Sirtuin-3 Inhibitor, LC-0296, Inhibits Cell Survival and Proliferation, and Promotes Apoptosis of Head and Neck Cancer Cells. Anticancer Res. 2016;36(1):49–60. [PMC free article] [PubMed] [Google Scholar]
  • 82.Napper AD, Hixon J, McDonagh T, Keavey K, Pons JF, Barker J, et al. Discovery of indoles as potent and selective inhibitors of the deacetylase SIRT1. J Med Chem. 2005;48(25):8045–8054. doi: 10.1021/jm050522v. [DOI] [PubMed] [Google Scholar]
  • 83.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. [DOI] [PubMed] [Google Scholar]
  • 84.Guan X, Lin P, Knoll E, Chakrabarti R. Correction: Mechanism of Inhibition of the Human Sirtuin Enzyme SIRT3 by Nicotinamide: Computational and Experimental Studies. PLoS One. 2015;10(8):e0136127. doi: 10.1371/journal.pone.0136127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Nguyen GT, Schaefer S, Gertz M, Weyand M, Steegborn C. Structures of human sirtuin 3 complexes with ADP-ribose and with carba-NAD+ and SRT1720: binding details and inhibition mechanism. Acta Crystallogr D Biol Crystallogr. 2013;69(Pt 8):1423–1432. doi: 10.1107/S0907444913015448. [DOI] [PubMed] [Google Scholar]
  • 86.Galli U, Mesenzani O, Coppo C, Sorba G, Canonico PL, Tron GC, et al. Identification of a sirtuin 3 inhibitor that displays selectivity over sirtuin 1 and 2. Eur J Med Chem. 2012;55:58–66. doi: 10.1016/j.ejmech.2012.07.001. [DOI] [PubMed] [Google Scholar]
  • 87.Disch JS, Evindar G, Chiu CH, Blum CA, Dai H, Jin L, et al. Discovery of thieno[3,2-d]pyrimidine-6-carboxamides as potent inhibitors of SIRT1, SIRT2, and SIRT3. J Med Chem. 2013;56(9):3666–3679. doi: 10.1021/jm400204k. [DOI] [PubMed] [Google Scholar]
  • 88.Patel K, Sherrill J, Mrksich M, Scholle MD. Discovery of SIRT3 Inhibitors Using SAMDI Mass Spectrometry. J Biomol Screen. 2015;20(7):842–848. doi: 10.1177/1087057115588512. [DOI] [PubMed] [Google Scholar]
  • 89.Lain S, Hollick JJ, Campbell J, Staples OD, Higgins M, Aoubala M, et al. Discovery, in vivo activity, and mechanism of action of a small-molecule p53 activator. Cancer Cell. 2008;13(5):454–463. doi: 10.1016/j.ccr.2008.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Gertz M, Nguyen GT, Fischer F, Suenkel B, Schlicker C, Franzel B, et al. A molecular mechanism for direct sirtuin activation by resveratrol. PLoS One. 2012;7(11):e49761. doi: 10.1371/journal.pone.0049761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Schuetz A, Min J, Antoshenko T, Wang CL, Allali-Hassani A, Dong A, et al. Structural basis of inhibition of the human NAD+-dependent deacetylase SIRT5 by suramin. Structure. 2007;15(3):377–389. doi: 10.1016/j.str.2007.02.002. [DOI] [PubMed] [Google Scholar]
  • 92.Zang W, Hao Y, Wang Z, Zheng W. Novel thiourea-based sirtuin inhibitory warheads. Bioorg Med Chem Lett. 2015;25(16):3319–3324. doi: 10.1016/j.bmcl.2015.05.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Suenkel B, Fischer F, Steegborn C. Inhibition of the human deacylase Sirtuin 5 by the indole GW5074. Bioorg Med Chem Lett. 2013;23(1):143–146. doi: 10.1016/j.bmcl.2012.10.136. [DOI] [PubMed] [Google Scholar]

RESOURCES