Open access peer-reviewed chapter

Circadian Regulation of Colon Cancer Stem Cells: Implications for Therapy

Written By

Sandra Ríos-Arrabal, José Antonio Muñoz-Gámez, Sergio Manuel Jiménez-Ruíz, Jorge Casado-Ruíz, Francisco Artacho-Cordón and Josefa León-López

Submitted: 28 September 2015 Reviewed: 29 January 2016 Published: 07 September 2016

DOI: 10.5772/62323

From the Edited Volume

Colorectal Cancer - From Pathogenesis to Treatment

Edited by Luis Rodrigo

Chapter metrics overview

1,775 Chapter Downloads

View Full Metrics

Abstract

The presence of cancer stem cells (CSCs) in colorectal cancer (CRC) has been associated with tumor initiation, metastasis, relapse, and resistance to chemotherapy and radiotherapy. Therefore, a better knowledge of the molecular mechanisms involved in the regulation of CSCs is required to develop treatments that are more effective. Like normal cells, cancer cells contain molecular clocks that generate circadian rhythms in gene expression and metabolic activity. Disruption of circadian rhythms has been linked to increased cancer risk, chemoresistance, and progression in CRC. CSCs also generate rhythms, which could be exploited with a chronopharmacological approach. Although the regulation of the expression of circadian rhythm genes appears to be mediated mainly by transcription–translation feedback loops, the existence of forms of nontranscriptional regulation has been demonstrated. Particularly, microRNAs (miRNA) and SIRT1 are significant players in regulating various aspects of the circadian clock function. Furthermore, miRNA acts as a regulator of cancer progression by regulating the CSC characteristics through SIRT1. These findings led us to hypothesize that there is a circadian rhythm of CSC markers regulated by miRNAs in CRC with SIRT1 acting as a mediator of miRNA activity. The pharmacological regulation of SIRT1, and therefore of the circadian machinery, could result in antiproliferative effects and increased sensitivity to antitumor treatments in CRC.

Keywords

  • circadian rhythms
  • cancer stem cells
  • SIRT1
  • redox homeostasis
  • melatonin

1. Introduction

Cancer is a major health burden and one of the leading causes of death worldwide. There are more than 200 known types of cancer, being CRC one of the most common cancers. In men, CRC is the third most common, after lung and prostate cancer. In women, CRC is the second most common cancer behind breast cancer [1].

Cancer begins with the transformation of a normal cell into a tumor cell through a multistage process. In this process, the changes are the result of the interaction between the genetic factors of the patient and external agents, including physical, chemical, or biological agents. Aging is another fundamental factor in the development of cancer. The coming years will see an increase in the total number of cancer cases [2]. This will be caused by the dramatic increase in average life expectancy during the last century in the industrialized world and by the fact that about 5% of the population in the developed countries is predicted to have more than 85 years by 2050 [3].

CRC is an excellent example of the increased cell malignancy with age; according to the Centers for Control and Prevention of Diseases, the incidence rate of CRC increases progressively with age. In 2007, the Surveillance, Epidemiology, and End Results Program registry reported that CRC incidence was 74.5/100,000 in persons aged 50–64 years, 186.0/100,000 in persons aged 65–74 years, and 290.1/100,000 in persons aged ≥75 years [4].

Little is known about the precise biochemical mechanisms responsible for the rise in CRC rates with aging. Some authors have proposed cancer stem cells/stem-like cells (CSCs/CSLCs) as the main factor associated with age-related rise in CRCs [5]. CSCs are a small subpopulation of cells that might play a critical role in CRC progression and resistance to chemotherapy and radiotherapy. Because of these resistances, CSCs play an important role in the processes of tumor recurrence and metastasis [68]. Therefore, it is increasingly likely that therapies that specifically target CSCs will be needed for the complete eradication of the tumor [7] because these cells are responsible for the morbidity and mortality associated with this type of cancer [9].

Advertisement

2. CSCs and CRC

CSCs, also known as tumor-initiating cells, were first identified by John Dick in acute myeloid leukemia in the late 1990s. Since then, they have been an intense focus of cancer research. CSCs have recently been identified in several solid tumors, including breast [10], colon [11], head and neck [12], lung [13], pancreas [14], and central nervous system cancer [15]. CSCs are defined as cancer cells that possess characteristics associated with normal stem cells, specifically the ability to self-renew and differentiate into multiple cell types, in particular into all the heterogeneous cell types found in a particular cancer sample [16] (Figure 1).

Figure 1.

The cancer stem cells (CSCs) theory suggests that tumors grow like normal tissues of the body. In this way, the CSCs are able to both self-renew extensively, copying to them, and to produce the tissue grow (cancer tissue). In this process, the CSCs produce transit amplifying cells that have the capacity to divide a certain number of times, and then, they can differentiate into specialized heterogeneous tumor cells that do not have the skill to divide and therefore do not contribute to tumor growth.

Little is known about the origin of CSCs. Several proposals have been made regarding the origin of CSCs and probably several answers are possible depending on the tumor type and the tumor phenotype. In brief, CSC can arise if there are mutations in a developing stem or progenitor cells, in adult stem cells or adult progenitor cells, or in differentiated cells that acquire stem-like attributes [17]. Some data pointed to adult stem cells as origin of CSC in CRC. In this way, according to the most accepted model of CRC progression [18], four to five mutations in tumor suppressor genes or oncogenes are required and this takes about 8–12 years. As the colonic mucosa is a highly dynamic tissue, the mucosal cells are constantly replaced with cells derived from crypt stem cells. Therefore, only the long-lived cells (stem cells) may serve as reservoirs for accumulation of such mutations and epigenetic modifications.

In CRC, the presence of CSC has been correlated to tumor initiation, progression, metastasis, relapse, and resistance to chemotherapy and radiotherapy. [19]. CSCs have been isolated from CRC and are identified by the expression of specific surface markers and by their functional characteristics: CD44, CD166, CD133, ALDH1 (aldehyde dehydrogenase isoform 1), and ESA (epithelial-specific antigen, also known as EpCAM) [20]. Moreover, CSCs have been involved in the transition from adenoma to carcinoma, and this issue is age-dependent. In addition, an age-related rise in CSC has been observed in normal colonic mucosa and in premalignant adenomatous polyps. Moreover, the CSCs found showed an increased expression of CD44, CD166, ALDH1, miR-21, and oncogenic miRNA, which suggests a predisposition of the organ to developing colorectal cancer [21]. Another characteristic of CSCs is that they can undergo epithelial–mesenchymal transition [22]. The fact that CSCs are resistant to conventional therapies has important implications in the development of novel strategies, such as CSCs-targeted treatments. It makes sense that eradication of CSCs or, alternatively, the attenuation of their malignant and stemness properties can lead us to achieve more effective therapeutic approaches. Therefore, development of specific CSCs-targeted therapies holds hope for improvement in the survival and quality of life of cancer patients, especially in patients with metastatic disease.

Advertisement

3. Epigenetic phenomena associated with CRC: miRNA regulatory network

Epigenetic processes are potentially reversible genetic modifications that lead to genomic instability and malfunction but that do not involve changes in DNA sequence. Epigenetic changes include (a) altered DNA methylation, (b) chromatin remodeling, and (c) small non-coding RNA (miRNAs). Notable changes in epigenetics have been reported for several age-related diseases, including CRC [23]. miRNAs are a small non-coding RNAs with 18–25 nucleotides found in plants, animals, and some viruses that are involved in RNA silencing and regulation of gene expression at the post-transcriptional level. miRNAs bind to the 3′-untranslated region (3′-UTR) of the protein-coding messenger RNAs (mRNAs). A single miRNA may regulate multiple mRNA targets and these small molecules are predicted to regulate approximately 60% of the human genes [24].

miRNAs are involved in different biological processes, including embryogenesis and the maintenance of stem cell characteristics, cell proliferation, metabolism, and transduction signals as in resistance to cancer treatments, including chemo- and radiotherapy [25, 26]. The presence of certain miRNAs has been associated with tumor development, tumor cell invasion, and cancer metastasis, and they may be of value as biomarkers for tumor detection and prognosis [27]. In addition, evidences show that miRNAs participate as oncogenic or tumor suppressors in the developmental and physiological processes of several human cancers, including CRC.

3.1. miRNA regulatory network in CRC

Approximately 450 unique miRNAs have been associated with CRC. In this disease, the aberrant expression of miRNAs has been associated with initiation and progression of CRC. Of special interest is miRNA involvement in the invasion, migration, and progression of disease through epithelial–mesenchymal transition into metastases. This event occurred due to IL6/STAT3 mediation and/or TP53 inactivation [28]. miRNAs are also involved in resistance to radio-chemotherapy [29].

Furthermore, miRNAs might regulate cell proliferation and apoptosis by targeting mitogen receptor tyrosine kinases and cell cycle regulators, such as cyclin-dependent kinases (CDKs), reciprocal miRNA interactions with MYC (v-myc avian myelocytomatosis viral oncogene homolog) and TP53, and by regulating the proapoptotic and antiapoptotic mRNAs [30].

Moreover, due to the altered expression of miRNAs in tumor development, they have been proposed as tissue and circulation biomarkers (blood derivates and feces). They might play a role in the prognostic and predictive diagnosis of CRC [31]. Table 1 shows the potential roles of miRNA in CRC. Additionally, miRNA-related polymorphisms might disrupt their binding sites, miRNA processing, and expression. In this way, the single-nucleotide polymorphism (SNP) and other genetic abnormalities might be of utility in the study of the risk and prognosis of CRC.

Role in CRC miRNA
Oncogene Proliferation miR-21, miR-92a, miR-96, miR-135a, miR-135b
Apoptosis miR-21
DNA damage response miR-155
Invasion miR-21, miR-92a
Epithelial–mesenchymal
transition
miR-92a
Metastasis miR-224
Inflammation miR-224
Tumor suppressor Proliferation let-7, miR-7, miR-18a, miR-26b, miR-27b, miR-143,
miR-144, miR-145, miR-194, miR-320a
Apoptosis miR-26b, miR-34a, miR-194, miR-195, miR-365, miR-491
Angiogenesis miR-27b, miR-145, miR-101, miR-126
Invasion miR-26b, miR-145, miR-194
Migration miR-26b, miR-145, miR-194
Circulatory biomarkers Diagnosis miR-21, miR-31, miR-34a, miR-92a, miR-143, miR-145, miR-601,
miR-760
Prognosis/survival miR-21, miR-124-5p, miR-141, miR-155, miR-182, miR-200c
Chemotherapy miR-17-5p, miR-19a, miR-20a, miR-27b, miR-126,
miR-130, miR-140, miR-145, miR-192, miR-216, miR-200c

Table 1.

MiRNAs described in CRC

MiRNA has also been described with high potential as therapeutic target. According to Schetter et al., the two strategies for miRNA-based therapies are (a) to inhibit oncogenic miRNAs and (b) to restore tumor suppressor miRNAs. Some preclinical models have shown that both strategies might be effective in the treatment of CRC cancers [29].

3.2. miRNA regulation of CSC in CRC

Many studies have reported that miRNAs are key players in the regulation of CSCs [32]. Bitarte et al. [33] have described that miR-451 is involved in the self-renewal, tumorigenicity, and chemoresistance of CRC stem cells. Up-regulation of miR-451 resulted in reduction of colon sphere formation and growth, inhibition of tumorigenicity in vivo, and sensitization to the active metabolite of irinotecan, SN38. This metabolite might be related to miR-451–mediated down-regulation of cyclooxygenase-2 (COX-2) and WNT (Wingless-type MMTV integration site family) pathway, essential to maintain cell stemness. On the other hand, Bitarte et al. also described that the regulation of expression of ATP-binding cassette (ABCB1) by miR-451 could decrease SN38 resistance.

Moreover, it has been suggested that miR-215 and miR-140 could control the slow proliferation and the chemoresistance of CSCs in the colon. In this respect, Jones et al. [34] described miRNA-215 (miR-215) as a direct transcriptional target of CDX1 in CRC stem cell differentiation. Song et al. [35] suggested miR-215 as molecular modulator of chemoresistance in CSC in CRC. Moreover, Yu et al. [36] have reported that miR-93 suppresses proliferation and colony formation of human CRC stem cells.

Advertisement

4. Circadian rhythms and CSC in CRC

Circadian rhythms are daily rhythms that take the form of a sine wave with high or active and low or quiet periods over the 24-hour clock. Many biological processes are temporally coordinated so that groups of genes called “clock genes” and their products are expressed at different critical times of the day, being likewise coordinated in circadian time [37]. A master clock in the suprachiasmatic nucleus (SCN) of the hypothalamus organizes the circadian system in a hierarchical manner. The SCN receives photic input through direct retinal innervation that initiates gene expression in the SCN [38]. In this way, light exposure synchronizes the SCN clock to solar time, adapting the oscillator to exact 24-hour cycle. The master clock of the SCN communicates day-night information to the rest of the body. The SCN receives light input from the retina and then conveys the photic information into neural or humoral signals. This information is then sent to peripheral circadian clocks that exist in almost all cells of the body and synchronize them to the same phase [39]. Whereas light is the dominant rhythmic signal for the SCN oscillator, the peripheral clocks respond to other environmental signal, such as body temperature, hunger, and hormone secretion cycles, and modify their phase accordingly [40, 41].

4.1. Regulation of circadian rhythms in mammals

In mammals, these daily rhythms are maintained by autoregulatory transcriptional and translational feedback and feed-forward loops (TTFLs) [42]. The core clock genes are BMAL1 (brain and muscle aryl-hydrocarbon receptor nuclear translocator-like 1), CLOCK (circadian locomotor output cycles kaput), PER (period homolog), and CRY (cryptochrome) [41]. BMAL1 heterodimerizes with either CLOCK or NPAS2 (neuronal PAS domain protein 2) and binds to E-box elements in PER (PER1-3) and CRY (CRY1-2) promoter regions and activates their transcription. Upon accumulation in the cytoplasm, PER and CRY proteins translocate to the nucleus where they repress the BMAL1: CLOCK/NPAS2 regulatory complex, thereby shutting down their own transcription. The PER and CRY heterodimers are progressively degraded, allowing the circuit to start again. This negative feedback leads to a cycle in gene expression that takes approximately 24 hours to complete (Figure 2) [43].

Figure 2.

Schematic representation of the core circadian clock. (Adapted from Robinson and Reddy [52]. Copyright ©2014). BMAL1 and CLOCK transcription products translocate to the cytoplasm and dimerize. They then return to the nucleus and bind to E-box regions and promote Period (PER) and Cryptochrome (CRY) transcription. PER and CRY translocate to the cytoplasm, dimerize, and return to the nucleus and inhibit the binding of the CLOCK/BMAL complex. PER and CRY are subsequently degraded. REV-ERB inhibits the transcription of Bmal1, whereas ROR promotes BMAL1 expression. The clock-controlled genes REV-ERBα (REV-ERB) and RORα (ROR) generate additional circadian control by modulating the expression of BMAL1.

Post-transcriptional and post-translational modifications, such as phosphorylation, acetylation, methylation, sumoylation, and ubiquitination, are crucial to the clock molecular mechanism [44]. Post-translational modifications of the proteins of the circuit generate the essential time delay that maintains the period of the cycle at approximately 24 hours. Additional feedback pathways involving nuclear receptors, such as RORα (retinoid-related orphan receptor alpha), REV-ERBα (NR1D1, nuclear receptor subfamily 1, group D, member 1), PPARγ (peroxisome proliferator–activated receptor gamma), and PGC-1α (peroxisome proliferator–activated receptor gamma coactivator 1-alpha) provide further robustness to the circuit [45, 46]. Additionally, these clock genes control numerous target genes (termed clock controlled genes, CCGs) and they regulate the circadian rhythms of various biochemical and physiological processes [47].

4.2. Redox regulation of circadian rhythms

The possible relation between circadian rhythm and redox state has been long known but it was unclear whether redox oscillations are a driver of the clock or a biomarker of cellular time. The discovery that circadian redox oscillations appear to be conserved throughout evolution, and that circadian oscillations in redox parameters persist in the absence of transcriptional system, has consolidated the interplay between metabolic processes and the molecular clock [48]. In mammals, nearly 10,000 genes are known to be under circadian control [49]. Several studies have investigated circadian oscillations in gene expression in metabolic tissues, including brown fat, liver, and skeletal muscle [50]. Certain metabolic enzymes are rhythmically expressed in the liver, such as aconitase, aldolase 2, enolase 1, ketohexokinase, and succinate dehydrogenase 1 [51]. Since numerous metabolic diseases appear to have a circadian-related dysfunction, it follows that core cellular metabolism and the clock are intimately connected [52].

The redox state of a cell has been shown to be an integral component in the regulation of the molecular clock. DNA binding of NPAS2 and CLOCK is influenced by the redox state of NAD(H) and NHAP(H). NADP inhibits NPAS2:BMAL1 binding to DNA, but, on the other hand, DNA binding is promoted by NADPH. At a ratio of NADP:NADP(H) of over 75% NPAS2:BMAL1 DNA binding increases, whereas below 75% binding decreases [53].

Peroxiredoxins (PRDXs) are a family of antioxidants that help to prevent cellular damage resulting from the production of ROS and work by reducing H2O2 to water. Six PRDX isoforms are known (PRDX 1–6), and they are found in different cellular localizations: isoforms 1, 2, 5, and 6 in the cytosol, isoforms 3 and 5 in the mitochondria, isoform 5 in the peroxisome, and isoform 4 in the endoplasmic reticulum [54, 55]. PRDXs could have one or two redox active cysteine residues that bind H2O2 forming a sulfenic acid. In mammals, there are five two-Cysperoxiredoxins (PRDX 1–5) and a single one-Cysperoxiredoxin (PRDX 6) [56]. In most cases, a homodimer is formed by a resolving cysteine at the C terminus forming a disulphide bond with the cysteine sulfenic acid. This bond can then be ‘resolved’ by thioredoxin, enabling further catabolism of peroxide by PRDX [57].

The cycle of PRDX had been shown to follow a circadian rhythm, with two forms of PRDX6 oscillating in antiphase in the liver, emphasizing a role of post-translational modification in circadian rhythmicity [51]. While the core clockwork use TTFLs, many works suggest the redox state of PRDXs is driven by the metabolic state of the cell, but it is independent of transcription [58]. However, the reciprocal influence between the TTFL and the PRDX-based metabolic clock has not fully addressed yet.

4.3. Epigenetic regulation of circadian rhythms by miRNAs

Some miRNAs have regulatory functions contributing to the control over circadian protein expression [59]. The influences of the circadian clock function on miRNA expression or vice versa have been well established. Experimental evidence suggests that up to 30% of core clock genes are under the control of miRNA. Recently, miR-132 and miR-219 seem to be directly involved in the clock system and regulate light response [60].

Other studies examine the role of extracellular miRNAs in the regulation of molecular components and modulation of rhythmicity in peripheral cellular clocks. The studies revealed that exposure to miR-142-3p– and miR-494–enriched conditioned medium increases intracellular expression of these miRNAs and results in functional effects in recipient cells [59].

4.4. Circadian rhythms and CRC

In carcinogenesis, the levels of expression of many proteins may be dependent directly or indirectly on the circadian cycle. Investigations on the relation between the circadian clock and DNA damage response have revealed that DNA damage checkpoints, nucleotide excision repair, and apoptosis are appreciably influenced by the clock [61]. Changes in the circadian rhythm of cell division are considered important component in neoplastic transformation. The presence of DNA damage is usually associated with cell cycle arrest for attempted repairs or induction of apoptosis. The mechanism of repair happens before the S-phase of the cell cycle, and although there are post-replication mechanisms for induction of cell cycle arrest, the G1/S checkpoint is usually the most stringent cell cycle checkpoint and this event will take place at night for adjustment of the circadian rhythm [61]. On the other hand, disruptions of the circadian rhythm genes are associated with increased susceptibility to cancers [62].

Some experimental studies have showed the role of the disruption of the molecular clock work in colorectal carcinogenesis and CRC progression [63]. One example is seen in human cancer cells lines when PER1 is overexpressed. In this experiment, it was observed a reduced colony formation and clonogenic expansion, in sensitization to radiation-induced apoptosis, and in altered expression of transcriptional target genes, such as c-MYC and p21 [64]. In contrast, PER2-null mice showed an increase in hyperplasia and neoplasia in response to γ-radiation [65]. In line with this work, restoring CLOCK expression in a human colon adenocarcinoma cell line derived from a primary colon cancer, in response to ionizing radiation, conferred protection against ultraviolet (UV)-induced apoptosis and decreased G2/M arrest [66].

On the other hand, one study that used CRC tumor tissue from patients demonstrated that PER2 expression was higher in well-differentiated cancer cells when compared to poorly differentiated ones. Associations of decreased PER2 levels with patient age, histological grade, TNM (for tumors/nodes/metastases) stage, and expression of nuclear proliferation-related antigen Ki67 were also observed [67]. In addition, down-regulation of PER3 associated with various clinicopathological factors, including tumor location, differentiation, and stage, as well as poorer survival was seen in CRC tissues, thus suggesting an important role in CRC progression [68].

The efficacy of many drugs and the intensity of their effects depend on the time of day when they are administered (chronochemotherapy), and they are therefore associated with the circadian rhythm. The clock acts as a modulator of the pharmacokinetics and pharmacodynamics of chemotherapeutic drugs and of the activity of the DNA-repair enzymes that repair the DNA damage caused by anticancer drugs [61]. Some studies have demonstrated that chronomodulated chemotherapy has better tolerability and antitumor activity compared with conventional chemotherapy. A good example is the treatment of advanced stage CRC where oxaliplatin is administered in the afternoon and 5-FU and leucovorin are delivered late at night (chronoFLO4) [69]. Other authors have observed that 5-FU administration during the sleeping time before radiotherapy could have an advantage as a chronotherapy and also as a radiosensitizer [70].

4.5. Regulation of stemness by the circadian machinery

Numerous studies are demonstrating the importance of coherent circadian oscillations for a variety of homeostatic functions of tissues. One example is the timed activation and differentiation of stem and progenitor cells, and how perturbation of this temporal coordination leads to pathologies, including obesity and neurological diseases, aging, and cancer [71].

Like normal cells, cancer cells contain molecular clocks that generate circadian rhythms in gene expression and metabolic activity [72]. The circadian rhythms not only regulate the cell cycle in normal differentiated cells but there could also be a functional relation between circadian rhythm gene expression and the intrinsic control of the proliferation of CSCs and progenitor cells in different tissues [73]. In rodents and humans, all the hematopoietic stem and progenitor cells exhibit a predictable circadian variation [74, 75]. Some authors have shown that there is an increase in circadian rhythm gene expression of highly differentiated stem cell cultures. In both cancer and normal cells, a rhythmic nuclear translocation of PER2 and other critical clock proteins is an essential part of a clock timing mechanism based on transcriptional–translational feedback loops and rhythmic chromatin modification [73, 76].

The malignant phenotype depends not only on the characteristics of the cancer cell itself but also on the tumor microenvironment. CSCs have to survive for a long time in the body to generate the highly tumorigenic cells responsible for the clinical manifestations of cancer. During this period, the niche helps to shelter CSCs from different types of insults, such as the immune response and chemotherapy-induced genotoxic stress [77, 78]. This suggests that the niche may also play a protective role for CSCs, thus increasing the risk of cancer. In fact, BMAL1 suppresses cancer cell invasion by blocking the phosphoinositide 3-kinase-Akt-Matrix metalloproteinase-2 (MMP-2) signaling pathway [79]. Other authors have reported circadian oscillations in the levels of MMP-9 [80].

Advertisement

5. SIRT1 and the circadian regulation of CSCs by miRNAs

The silent information regulator (SIRT) 2 family of proteins, known as sirtuins, is a class III of histone deacetylases or NAD+−dependent deacetylases and is conserved from bacteria to humans. The requirement for NAD+ links the activity of sirtuins directly to the metabolic state of the cell, since the deacetylase activity of these proteins is controlled by the cellular NAD+/NADH ratio. There are seven sirtuins (SIRT1–7) in mammals. They have different specific substrates and biological functions and are found in various cell compartments [81].

The role of sirtuin activation in mammals is to regulate the progression of aging and age-associated disorders, including neurodegeneration, diabetes, cardiovascular diseases, and many types of cancer [82]. The best characterized and well-studied among the human sirtuins is SIRT1. It can be found in the nucleus and in the cytoplasm of the cells [83].

5.1. Sirt1 in cancer biology

SIRT1 is overexpressed in some types of human cancer tissues, such as ovary, liver, breast, stomach, pancreas and prostate, and down expressed in skin cancer. In CRC, SIRT1 is also overexpressed. However, other investigations have revealed pronounced SIRT1 expression in both normal colon and tumor tissues, although its expression is substantially reduced in higher grade CRC tumors [84].

SIRT1 has a dual role in tumorigenesis, where it can function as either a tumor promoter or a tumor suppressor. Its function in malignancy varies with concentration, cellular location, temporal and spatial distribution, and regulation by upstream and downstream factors [85].

The initial connection of sirtuins to cancer was made when SIRT1 was found to deacetylate and repress the activity of the tumor suppressor p53 [86, 87]. SIRT1-mediated deacetylation suppresses the functions of other tumor suppressors, including p73, hypermethylated in cancer 1 (HIC1), E2F transcription factor 1(E2F1, also known as retinoblastoma-associated protein 1), retinoblastoma protein (Rb), and phosphatase and tensin homologue deleted in chromosome 10 (PTEN), thus suggesting that SIRT1 acts as a promoter in tumor development and progression [8893]. Other reports showed SIRT1 as a downstream of the oncoprotein BCR-ABL tyrosine kinase (Abelson murine leukemia viral oncogene homolog 1), implicated in the development of chronic myelogenous leukemia (CML)-like myeloproliferative disease [94].

Autophagy is a self-degradative process that plays a role by eliminating damaged organelles and misfolded or aggregated proteins through the lysosomal degradation pathway. Autophagy initially serves as a protective process to prevent cancer initiation; however, after neoplastic transformation, it can promote tumor cell survival and maintenance [95]. Autophagy can also affect chemotherapeutic and immunotherapeutic response in cancer cells making it an attractive target for development of anticancer drugs [96, 97]. SIRT1 forms a molecular complex with the genes related to autophagy and autophagosome formation, Atg5, Atg7, and Atg8. Loss of SIRT1 activity results in the acetylation of those factors thus leading to defects in the process of autophagy [98].

Consistent with a tumor-suppressor role, SIRT1 deacetylates and decreases the stability of the oncogene c-MYC [99]. In addition, the activity of SIRT1 can be increased by some tumor-suppressors, for example BRCA1 (breast cancer 1) [99]. SIRT1 exerts anticarcinogenic effects through multiple mechanisms [83]. SIRT1 can counteract various genotoxic insults, including oxidative DNA damage, thereby blocking initiation of carcinogenesis. SIRT1 deacetylates and inhibits proapoptotic p53 and PARP-1 (poly (ADP-ribose) polymerase 1) under stressful conditions, conferring cell survival [86, 100]. SIRT1 is also required for DNA repair processes to maintain genomic stability [101, 102].

SIRT1 inhibits the mediators involved in aberrantly amplified proinflammatory signaling during promotion and progression of carcinogenesis [103105]. The anti-inflammatory effect of SIRT1 might be achieved by inhibition of several transcription factors related to inflammation, nuclear factor κB (NF-κB), signal transducer and activator of transcription 3 (STAT3), and the c-Jun and fos elements of transcription factor activator protein 1 (AP-1) [106108]. Mainly through NF-κB and AP-1 pathways, SIRT1 was engaged in macrophage and T-cell activation [105, 109]. SIRT1 also regulates the differentiation and function of iTreg (induced Treg helper cells) [110]. Interestingly, SIRT1 translates metabolic cues during regulation of the immune responses, which would bring new insights into both pathogenesis and potential therapeutic strategies of a variety of immune-related diseases, such as cancer [111].

5.2. SIRT1 and stem cells

SIRT1 is considered an old multifaceted enzyme with an important role in the maintenance of pluripotency in various types of stem cells. Most of the in vivo data suggest that Sirt1 acts in early development as a modulatory molecule on basic developmental processes [112]. In cancer, SIRT1 has been implicated in the regulation of CSCs survival and differentiation. SIRT1 has been found to regulate the growth and survival of leukemia stem cells (LSCs) and confer resistance against chemotherapy [113], stimulate endometrial cell tumor growth through lipogenesis [114], maintain neural stem cells and promote oncogenic transformation [115], and foster hepatocellular carcinoma [116]. As a result, SIRT1 and agents that modulate SIRT1 activity may represent new therapeutic strategies against tumorigenesis.

One of the more intriguing hypotheses about aging and age-related disease is that age-associated phenotypic alterations derive from the inability of resident stem cells to maintain tissue structure and function [117]. This suggests that the aging process could arise from loss or malfunction of self-renewal and/or differentiation potential in adult stem cell populations. SIRT1 has a positive role in stemness by aiding in the silencing of differentiation genes, which suggests new potential explanations of its ability to extend lifespan and to avoid cell and organism senescence [112, 118].

5.3. SIRT1 and miRNAs

Expression of SIRT1 is controlled at multiple levels by transcriptional, post-transcriptional, and post-translational mechanisms under physiological and pathological conditions [119]. Deacetylation activity of SIRT1 can be modulated by multiple regulators. AROS (Active regulator of SIRT1) and DBC1 (deleted in breast cancer 1) are positive and negative regulators of SIRT1, respectively [120, 121].

Emerging evidence indicates that miRNAs are important regulators of SIRT1 expression and activity [119]. In cancer, SIRT1 mediates miR-34a activation of apoptosis by regulating p53 activity. In addition, p53 induces expression of miR-34a which suppresses SIRT1, increasing p53 activity [122]. In CRC, dysregulation of microRNA-34a expression causes drug-resistance to 5-FU in human colon cancer cells through the downregulation of Sirt1 and E2F3 [123, 124].

miRNAs play an important role in proper function and differentiation of human and mouse stem cells. Recently, it has been demonstrated that miR-34a is required for proper differentiation of mouse embryonic stem cells, mouse neural stem cell, and mouse embryonic fibroblasts and that it function in part by targeting SIRT1 and modulating p53 activity [125127]. Also, the miR-29b-Sirt1 axis regulates self-renewal of mouse embryonic stem cells in response to reactive oxygen species [128]. miR-34a is also a critical regulator of cancer progression by the regulation of CSC characteristics, through SIRT1 as a mediator [129133], and mainly, through up-regulation of p53/p21 [131].

5.4. SIRT1 as regulator of circadian rhythms

Two independent studies identified SIRT1 as a critical modulator of the circadian clock machinery. Asher et al. [134] observed oscillations in SIRT1 protein levels, and Nakahata et al. [135] demonstrated that SIRT1 activity, and not its protein levels, oscillates in a circadian manner. SIRT1 modulates circadian rhythms by deacetylating histone H3 Lys9 and Lys14 at promoters of rhythmic genes, BMAL1 and PER2. The CLOCK-BMAL1 complex interacts with SIRT1 and recruits it to the promoters of rhythmic genes. While BMAL1 acetylation acts as a signal for CRY recruitment, PER2 acetylation enhances its stability [134]. These findings led to the concept that SIRT1 operates as a rheostat of the circadian machinery, modulating the amplitude of CLOCK-mediated acetylation and consequent transcription cycles [135].

Circadian oscillation of SIRT1 activity suggested that cellular NAD+ levels may also oscillate. In fact, circadian clock controls the expression of NAMPT (nicotinamide phosphoribosyltransferase), a key rate-limiting enzyme in the salvage pathway of NAD+ biosynthesis. The oscillatory expression of NAMPT is abolished in Clock/Clock mice, which results in drastically reduced levels of NAD+. These results imply the existence of an enzymatic/transcriptional feedback loop, wherein SIRT1 regulates the levels of its own cofactor [135]. These results also connect the circadian machinery to cell metabolism [134].

SIRT1 either directly or indirectly can influence the redox property of the cell [136]. In addition to reduce cellular oxidative stress burden, SIRT1 is also regulated by oxidative stress [137]. Since redox homeostasis influence circadian machinery, SIR1 could regulate circadian genes trough redox status of the cell. A recent report suggests that PRDX2 regulates the TTFL oscillation by decreasing the nuclear redox levels and increasing SIRT1 enzymatic activity, although neither a direct interaction between PRX and SIRT1 nor a modulation of SIRT1 intracellular levels by PRX was found [138].

5.5. SIRT1 regulators as new tools for cancer treatment

Recently, multiple research groups have pursued the identification and development of small molecule compounds that modulate sirtuins SIRT1 regulators as new tools for cancer treatment [139]. To date, SIRT1 inhibitors and activators have been described with different effects on cancer [140].

SIRT inhibitors require combined targeting of both SIRT1 and SIRT2 to induce p53 acetylation and cell death, like sirtinol and salermide [139]. Trichostatin A (TSA) and sirtinol induce p38MAPK- and AMPK-mediated downregulation of survivin and its functional correlation with decreased colon cancer cell viability in vitro [141]. Other salermide-related sirtuin inhibitors show antiproliferative effects in colon cancer cells in vitro, including CSCs [142]. Vorinostat activates p53, but does not require p53 for inducing its anticancer action in CRC [143]. The SIRT1 selective inhibitor Amurensin G may be effective in eliminating colon CSCs and may be applicable to potentiate the sensitivity of colon CSCs to TNF-related apoptosis-inducing ligand (TRAIL) [144].

Studies have shown that the sirtuin activator resveratrol, a polyphenol found in wines has chemopreventive activity against various cancers. In CRC, resveratrol induce apoptosis and suppressed the PI3K/Akt signaling pathway. The combination treatment with resveratrol and 5-FU induced a synergistic enhancement of growth inhibition and apoptosis in colon cancer DLD-1 cells. Interestingly, resveratrol increased the intracellular expression level of miR-34a, which down-regulated the target gene E2F3 and its downstream Sirt1, resulting in growth inhibition [145].

Melatonin is the main secretory product of the pineal gland and plays important roles in several biological functions, including circadian rhythms, sleep, mood, reproductive physiology, and aging diseases [146]. Numerous studies based on animal and clinical data have provided evidence that melatonin reduces the incidence of experimentally induced cancers and may significantly inhibit the growth of some human tumors [147]. Recently, melatonin was confirmed as a novel inhibitor of SIRT1. Melatonin inhibits prostate cancer and osteosarcoma cell growth through SIRT1 inhibition [148, 149]. Interestingly, it was recently reported that melatonin decreases CSCs and dysplastic injuries in colon tissue [150].

Advertisement

6. Concluding remarks

SIRT1 expression correlated with depth of invasion, lymph node metastasis, and TNM stage in CRC. Simultaneously, SIRT1 overexpression predicted a poor overall survival in CRC patients, and SIRT1 is a candidate negative prognostic biomarker for CRC patients [151]. SIRT1 has also been implicated in chemoresistance in CRC patients with [152] or without metastasis [124]. Further investigations aimed at targeting SIRT1 alone or in combination with chemotherapy deserve further attention and may ultimately increase response rates in the treatment of CRC. In this sense, melatonin can significantly amplify the cytostatic and the cytotoxic effects triggered by other compounds or conventional drugs. We are far from having a satisfactory understanding about how and when melatonin exerts its beneficial effects. Melatonin in the nanomolar range induces down-regulation of SIRT1. This finding is of great relevance because there is intense research ongoing to identify nontoxic feasible inhibitors of SIRT1. Melatonin should be evaluated for the management of those cancers where this protein is overexpressed and is functional.

References

  1. 1. Ferlay J, Soerjomataram I, Dikshit R, Eser S, Mathers C, Rebelo M, Parkin DM, Forman D, Bray F. Cancer incidence and mortality worldwide: sources, methods and major patterns in GLOBOCAN 2012. International Journal of Cancer. 2015;136(5):E359–86. DOI: 10.1002/ijc.29210
  2. 2. Vaiopoulos AG, Kostakis ID, Koutsilieris M, Papavassiliou AG. Colorectal cancer stem cells. Stem Cells. 2012;30(3):363–71. DOI: 10.1002/stem.1031
  3. 3. Todaro M, Francipane MG, Medema JP, Stassi G. Colon cancer stem cells: promise of targeted therapy. Gastroenterology. 2010;138(6):2151–62. DOI: 10.1053/j.gastro.2009.12.063
  4. 4. Kemper K, Grandela C, Medema JP. Molecular identification and targeting of colorectal cancer stem cells. Oncotarget. 2010;1(6):387–95.
  5. 5. Nangia-Makker P, Yu Y, Majumdar AP. Role of cancer stem cells in age-related rise in colorectal cancer. World Journal of Gastrointestinal Pathophysiology. 2015;6(4):86–9. DOI: 10.4291/wjgp.v6.i4.86
  6. 6. Li X, Lewis MT, Huang J, Gutierrez C, Osborne CK, Wu MF, Hilsenbeck SG, Pavlick A, Zhang X, Chamness GC, Wong H, Rosen J, Chang JC. Intrinsic resistance of tumorigenic breast cancer cells to chemotherapy. Journal of the National Cancer Institute. 2008;100(9):672–9. DOI: 10.1093/jnci/djn123
  7. 7. Ogawa K, Yoshioka Y, Isohashi F, Seo Y, Yoshida K, Yamazaki H. Radiotherapy targeting cancer stem cells: current views and future perspectives. Anticancer Research. 2013;33(3):747–54.
  8. 8. Meirelles K, Benedict LA, Dombkowski D, Pepin D, Preffer FI, Teixeira J, Tanwar PS, Young RH, MacLaughlin DT, Donahoe PK, Wei X. Human ovarian cancer stem/progenitor cells are stimulated by doxorubicin but inhibited by Mullerian inhibiting substance. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(7):2358–63. DOI: 10.1073/pnas.1120733109
  9. 9. Nigam A. Breast cancer stem cells, pathways and therapeutic perspectives 2011. The Indian Journal of surgery. 2013;75(3):170–80. DOI: 10.1007/s12262-012-0616-3
  10. 10. Al-Hajj M, Wicha MS, Benito-Hernandez A, Morrison SJ, Clarke MF. Prospective identification of tumorigenic breast cancer cells. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(7):3983–8. DOI: 10.1073/pnas.0530291100
  11. 11. Dalerba P, Dylla SJ, Park IK, Liu R, Wang X, Cho RW, Hoey T, Gurney A, Huang EH, Simeone DM, Shelton AA, Parmiani G, Castelli C, Clarke MF. Phenotypic characterization of human colorectal cancer stem cells. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(24):10158–63. DOI: 10.1073/pnas.0703478104
  12. 12. Prince ME, Sivanandan R, Kaczorowski A, Wolf GT, Kaplan MJ, Dalerba P, Weissman IL, Clarke MF, Ailles LE. Identification of a subpopulation of cells with cancer stem cell properties in head and neck squamous cell carcinoma. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(3):973–8. DOI: 10.1073/pnas.0610117104
  13. 13. Ho MM, Ng AV, Lam S, Hung JY. Side population in human lung cancer cell lines and tumors is enriched with stem-like cancer cells. Cancer Research. 2007;67(10):4827–33. DOI: 10.1158/0008-5472.CAN-06-3557
  14. 14. Li C, Lee CJ, Simeone DM. Identification of human pancreatic cancer stem cells. Methods in Molecular Biology. 2009;568:161–73. DOI: 10.1007/978-1-59745-280-9_10
  15. 15. Singh SK, Hawkins C, Clarke ID, Squire JA, Bayani J, Hide T, Henkelman RM, Cusimano MD, Dirks PB. Identification of human brain tumour initiating cells. Nature. 2004;432(7015):396–401. DOI: 10.1038/nature03128
  16. 16. Clarke MF, Dick JE, Dirks PB, Eaves CJ, Jamieson CH, Jones DL, Visvader J, Weissman IL, Wahl GM. Cancer stem cells--perspectives on current status and future directions: AACR Workshop on cancer stem cells. Cancer Research. 2006;66(19):9339–44. DOI: 10.1158/0008-5472.CAN-06-3126
  17. 17. Rapp UR, Ceteci F, Schreck R. Oncogene-induced plasticity and cancer stem cells. Cell Cycle. 2008;7(1):45–51.
  18. 18. Fearon ER, Vogelstein B. A genetic model for colorectal tumorigenesis. Cell. 1990;61(5):759–67.
  19. 19. Roy S, Majumdar AP. Cancer stem cells in colorectal cancer: genetic and epigenetic changes. Journal of Stem Cell Research & Therapy. 2012;Suppl 7(6). DOI: 31 10.4172/2157-7633.S7-006
  20. 20. Rassouli FB, Matin MM, Saeinasab M. Cancer stem cells in human digestive tract malignancies. Tumour Biology: The Journal of the International Society for Oncodevelopmental Biology and Medicine. 2015. DOI: 10.1007/s13277-015-4155-y. [Epub ahead of print].
  21. 21. Nautiyal J, Du J, Yu Y, Kanwar SS, Levi E, Majumdar AP. EGFR regulation of colon cancer stem-like cells during aging and in response to the colonic carcinogen dimethylhydrazine American Journal of Physiology. Gastrointestinal and Liver Physiology. 2012;302(7):G655-63. DOI: 10.1152/ajpgi.00323.2011
  22. 22. Kanwar SS, Yu Y, Nautiyal J, Patel BB, Majumdar AP. The Wnt/beta-catenin pathway regulates growth and maintenance of colonospheres. Molecular Cancer. 2010;9:212. DOI: 10.1186/1476-4598-9-212
  23. 23. Liu L, Wylie RC, Andrews LG, Tollefsbol TO. Aging, cancer and nutrition: the DNA methylation connection. Mechanisms of Ageing and Development. 2003;124(10–12):989–98.
  24. 24. Friedman RC, Farh KK, Burge CB, Bartel DP. Most mammalian mRNAs are conserved targets of microRNAs. Genome Research. 2009;19(1):92–105. DOI: 10.1101/gr.082701.108
  25. 25. Nam EJ, Yoon H, Kim SW, Kim H, Kim YT, Kim JH, Kim JW, Kim S. MicroRNA expression profiles in serous ovarian carcinoma. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research. 2008;14(9):2690–5. DOI: 10.1158/1078-0432.CRC-07-1731
  26. 26. Pourrajab F, Babaei Zarch M, BaghiYazdi M, Hekmatimoghaddam S, Zare-Khormizi MR. MicroRNA-based system in stem cell reprogramming; differentiation/dedifferentiation. The International Journal of Biochemistry & Cell Biology. 2014;55:318–28. DOI: 10.1016/j.biocel.2014.08.008
  27. 27. Malek E, Jagannathan S, Driscoll JJ. Correlation of long non-coding RNA expression with metastasis, drug resistance and clinical outcome in cancer. Oncotarget. 2014;5(18):8027–38.
  28. 28. Rokavec M, Oner MG, Li H, Jackstadt R, Jiang L, Lodygin D, Kaller M, Horst D, Ziegler PK, Schwitalla S, Slotta-Huspenina J, Bader FG, Greten FR, Hermeking H. IL-6R/STAT3/miR-34a feedback loop promotes EMT-mediated colorectal cancer invasion and metastasis. The Journal of Clinical Investigation. 2014;124(4):1853–67. DOI: 10.1172/JCI73531
  29. 29. Schetter AJ, Okayama H, Harris CC. The role of microRNAs in colorectal cancer. Cancer Journal. 2012;18(3):244–52. DOI: 10.1097/PPO.0b013e318258b78f
  30. 30. Cekaite L, Eide PW, Lind GE, Skotheim RI, Lothe RA. MicroRNAs as growth regulators, their function and biomarker status in colorectal cancer. Oncotarget. 2015. DOI: 10.18632/oncotarget.6390. [Epub ahead of print].
  31. 31. Meiri E, Mueller WC, Rosenwald S, Zepeniuk M, Klinke E, Edmonston TB, Werner M, Lass U, Barshack I, Feinmesser M, Huszar M, Fogt F, Ashkenazi K, Sanden M, Goren E, Dromi N, Zion O, Burnstein I, Chajut A, Spector Y, Aharonov R. A second-generation microRNA-based assay for diagnosing tumor tissue origin. The Oncologist. 2012;17(6):801–12. DOI: 10.1634/theoncologist.2011-0466
  32. 32. Liu X, Fu Q, Du Y, Yang Y, Cho WC. MicroRNA as regulators of cancer stem cells and chemoresistance in colorectal cancer. Current Cancer Drug Targets. 2015. [Epub ahead of print].
  33. 33. Bitarte N, Bandres E, Boni V, Zarate R, Rodriguez J, Gonzalez-Huarriz M, Lopez I, Javier Sola J, Alonso MM, Fortes P, Garcia-Foncillas J. MicroRNA-451 is involved in the self-renewal, tumorigenicity, and chemoresistance of colorectal cancer stem cells. Stem Cells. 2011;29(11):1661–71. DOI: 10.1002/stem.741
  34. 34. Jones MF, Hara T, Francis P, Li XL, Bilke S, Zhu Y, Pineda M, Subramanian M, Bodmer WF, Lal A. The CDX1-microRNA-215 axis regulates colorectal cancer stem cell differentiation. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(13):E1550-8. DOI: 10.1073/pnas.1503370112
  35. 35. Song B, Wang Y, Titmus MA, Botchkina G, Formentini A, Kornmann M, Ju J. Molecular mechanism of chemoresistance by miR-215 in osteosarcoma and colon cancer cells. Molecular Cancer. 2010;9:96. DOI: 10.1186/1476-4598-9-96
  36. 36. Yu XF, Zou J, Bao ZJ, Dong J. miR-93 suppresses proliferation and colony formation of human colon cancer stem cells. World Journal of Gastroenterology. 2011;17(42):4711–7. DOI: 10.3748/wjg.v17.i42.4711
  37. 37. Hrushesky W, Rich IN. Measuring stem cell circadian rhythm. Methods in Molecular Biology. 2015;1235:81-95. doi: 10.1007/978-1-4939-1785-3_8.
  38. 38. Hastings MH, Herzog ED. Clock genes, oscillators, and cellular networks in the suprachiasmatic nuclei. Journal of Biological Rhythms. 2004;19(5):400–13. DOI: 10.1177/0748730404268786
  39. 39. Mohawk JA, Green CB, Takahashi JS. Central and peripheral circadian clocks in mammals. Annual Review of Neuroscience. 2012;35:445. DOI: 10.1146/annurev-neuro-060909-153128
  40. 40. Kohsaka A, Waki H, Cui H, Gouraud SS, Maeda M. Integration of metabolic and cardiovascular diurnal rhythms by circadian clock. Endocrine Journal. 2012;59(6):447–56. DOI: 10.1507/endocrj
  41. 41. Chen L, Yang G. Recent advances in circadian rhythms in cardiovascular system. Frontiers in Pharmacology. 2015;6. DOI: 10.3389/fphar.2015.00071
  42. 42. Yang G, Paschos G, Curtis AM, Musiek ES, McLoughlin SC, FitzGerald GA. Knitting up the raveled sleave of care. Science Translational Medicine. 2013;5(212):212rv3. DOI: 10.1126/scitranslmed.3007225
  43. 43. Ukai H, Ueda HR. Systems biology of mammalian circadian clocks. Annual Review of Physiology. 2010;72:579–603. DOI: 10.1146/annurev-physiol-073109-130051
  44. 44. Cardone L, Hirayama J, Giordano F, Tamaru T, Palvimo JJ, Sassone-Corsi P. Circadian clock control by SUMOylation of BMAL1. Science. 2005;309(5739):1390–4. DOI: 10.1126/science.1110689
  45. 45. Liu C, Li S, Liu T, Borjigin J, Lin JD. Transcriptional coactivator PGC-1 alpha integrates the mammalian clock and energy metabolism. Nature. 2007;447(7143):477–81. DOI: 10.1038/nature05767
  46. 46. Yang G, Jia Z, Aoyagi T, McClain D, Mortensen RM, Yang T. Systemic PPAR gamma deletion impairs circadian rhythms of behavior and metabolism. Plos One. 2012;7(8):e38117 DOI:10.1177/1534735409352083
  47. 47. Paschos G. Circadian clocks, feeding time and metabolic homeostasis. Name: Frontiers in Pharmacology. 2015;6:112. DOI: 10.3389/fphar.2015.00112
  48. 48. Edgar RS, Green EW, Zhao Y, van Ooijen G, Olmedo M, Qin X, Xu Y, Pan M, Valekunja UK, Feeney KA. Peroxiredoxins are conserved markers of circadian rhythms. Nature. 2012;485(7399):459–64. DOI: 10.1038/nature11088
  49. 49. Ueda HR, Chen W, Adachi A, Wakamatsu H, Hayashi S, Takasugi T, Nagano M, Nakahama K, Suzuki Y, Sugano S. A transcription factor response element for gene expression during circadian night. Nature. 2002;418(6897):534–9. DOI: 10.1038/nature00906
  50. 50. Storch K-F, Lipan O, Leykin I, Viswanathan N, Davis FC, Wong WH, Weitz CJ. Extensive and divergent circadian gene expression in liver and heart. Nature. 2002;417(6884):78–83. DOI: 10.1038/nature744
  51. 51. Reddy AB, Karp NA, Maywood ES, Sage EA, Deery M, O'Neill JS, Wong GK, Chesham J, Odell M, Lilley KS. Circadian orchestration of the hepatic proteome. Current Biology. 2006;16(11):1107–15. DOI: 10.1016/j.cub.2006.04.026
  52. 52. Robinson I, Reddy A. Molecular mechanisms of the circadian clockwork in mammals. FEBS Letters. 2014;588(15):2477–83. DOI: 10.1016/j.febslet.2014.06.005
  53. 53. Rutter J, Reick M, Wu LC, McKnight SL. Regulation of clock and NPAS2 DNA binding by the redox state of NAD cofactors. Science. 2001;293(5529):510–4. DOI: 10.1126/science.1060698
  54. 54. Rhee SG, Woo HA. Multiple functions of peroxiredoxins: peroxidases, sensors and regulators of the intracellular messenger H2O2, and protein chaperones. Antioxidants & Redox Signaling. 2011;15(3):781–94. DOI: 10.1089/ars.2010.3393
  55. 55. Rhee SG, Woo HA, Kil IS, Bae SH. Peroxiredoxin functions as a peroxidase and a regulator and sensor of local peroxides. Journal of Biological Chemistry. 2012;287(7):4403–10. DOI: 10.1074/jbc.R111.283432
  56. 56. Hall A, Karplus PA, Poole LB. Typical 2-Cys peroxiredoxins: structures, mechanisms and functions. FEBS Journal. 2009;276(9):2469–77. DOI: 10.1111/j.1742-4658.2009.06985.x.
  57. 57. Hoyle NP, O'Neill JS. Oxidation-reduction cycles of peroxiredoxin proteins and nontranscriptional aspects of timekeeping. Biochemistry. 2015;54(2):184–93. DOI: 10.1021/bi5008386
  58. 58. Avitabile D, Ranieri D, Nicolussi A, D'Inzeo S, Capriotti AL, Genovese L, Proietti S, Cucina A, Coppa A, Samperi R. Peroxiredoxin 2 nuclear levels are regulated by circadian clock synchronization in human keratinocytes. The International Journal of Biochemistry & Cell Biology. 2014;53:24–34. DOI: 10.1016/j.biocel.2014.04.024
  59. 59. Shende VR, Kim S-M, Neuendorff N, Earnest DJ. MicroRNAs function as cis-and trans-acting modulators of peripheral circadian clocks. FEBS Letters. 2014;588(17):3015–22. DOI: 10.1016/j.febslet.2014.05.058
  60. 60. Cheng H-YM, Papp JW, Varlamova O, Dziema H, Russell B, Curfman JP, Nakazawa T, Shimizu K, Okamura H, Impey S. microRNA modulation of circadian-clock period and entrainment. Neuron. 2007;54(5):813–29. DOI: 10.1016/j.neuron.2007.05.017
  61. 61. Sancar A, Lindsey-Boltz LA, Gaddameedhi S, Selby CP, Ye R, Chiou Y-Y, Kemp MG, Hu J, Lee JH, Ozturk N. Circadian clock, cancer, and chemotherapy. Biochemistry. 2015;54(2):110–23. DOI: 10.1021/bi5007354
  62. 62. Kettner NM, Katchy CA, Fu L. Circadian gene variants in cancer. Annals of Medicine. 2014;46(4):208–20. DOI: 10.3109/07853890.2014.914808
  63. 63. Mazzoccoli G, Vinciguerra M, Papa G, Piepoli A. Circadian clock circuitry in colorectal cancer. World Journal of Gastroenterology. 2014;20(15):4197. DOI: 10.1016/jbbrc.2006.05.094
  64. 64. Yang X, Wood PA, Ansell C, Hrushesky WJ. Circadian time-dependent tumor suppressor function of period genes. Integrative Cancer Therapies. 2009;8(4):309–16. DOI: 10.1177/1534735409352083
  65. 65. Wood PA, Yang X, Hrushesky WJ. Clock genes and cancer. Integrative Cancer Therapies. 2009;8(4):303–8. DOI: 10.1177/1534735409355292
  66. 66. Alhopuro P, Björklund M, Sammalkorpi H, Turunen M, Tuupanen S, Biström M, Niittymäki I, Lehtonen HJ, Kivioja T, Launonen V. Mutations in the circadian gene CLOCK in colorectal cancer. Molecular Cancer Research. 2010;8(7):952–60. DOI: 10.1158/1541-7786.MCR-10-0086
  67. 67. Wang Y, Hua L, Lu C, Chen Z. Expression of circadian clock gene human Period2 (hPer2) in human colorectal carcinoma. World Journal of Surgical Oncology. 2011;9(1):166. DOI: 10.1186/1477-7819-9-166
  68. 68. Wang X, Yan D, Teng M, Fan J, Zhou C, Li D, Qiu G, Sun X, Li T, Xing T. Reduced expression of PER3 is associated with incidence and development of colon cancer. Annals of Surgical Oncology. 2012;19(9):3081–8. DOI: 10.1245/s10434-012-2279-5
  69. 69. Levi F, Schibler U. Circadian rhythms: mechanisms and therapeutic implications. Annual Review of Pharmacology and Toxicology. 2007;47:593–628. DOI: 10.1146/annurev.pharmtox.47.120505.105208
  70. 70. Asao T, Sakurai H, Harashima K, Yamaguchi S, Tsutsumi S, Nonaka T, Shioya M, Nakano T, Kuwano H. The synchronization of chemotherapy to circadian rhythms and irradiation in pre-operative chemoradiation therapy with hyperthermia for local advanced rectal cancer. Abbreviations: 5-FU 5-fluorouracil LV Leucovorin APR abdomino-perineal resection. International Journal of Hyperthermia. 2006;22(5):399–406. DOI: 10.1080/02656730600799873
  71. 71. Janich P, Meng QJ, Benitah SA. Circadian control of tissue homeostasis and adult stem cells. Current Opinion in Cell Biology. 2014;31:8–15. DOI: 10.1016/j.ceb.2014.06.010
  72. 72. Fujioka A, Takashima N, Shigeyoshi Y. Circadian rhythm generation in a glioma cell line. Biochemical and Biophysical Research Communications. 2006;346(1):169–74. DOI: 10.1016/jbbrc.2006.05.094
  73. 73. Sharma VP, Anderson NT, Geusz ME. Circadian properties of cancer stem cells in glioma cell cultures and tumorspheres. Cancer Letters. 2014;345(1):65–74. DOI: 10.1016/j.canlet.2013.11.009
  74. 74. Laerum O. Hematopoiesis occurs in rhythms. Experimental Hematology. 1995;23(11):1145–7. DOI: 10.1111/j.1600-0609.1997.tb01680.x
  75. 75. Bourin P, Ledain AF, Beau J, Mille D, Lévi F. In-vitro circadian rhythm of murine bone marrow progenitor production. Chronobiology International. 2002;19(1):57–67. DOI: 10.1081/CBI-120002677
  76. 76. Lowrey PL, Takahashi JS. Genetics of circadian rhythms in mammalian model organisms. Advances in Genetics. 2011;74:175. DOI: 10.1016/B978-0-12-387690-4.00006-4
  77. 77. Li L, Xie T. Stem cell niche: structure and function. Annual Review of Cell and Developmental Biology. 2005;21:605–31. DOI: 10.1146/annurev.cellbio.21.012704.131525
  78. 78. McAllister SS, Weinberg RA. Tumor-host interactions: a far-reaching relationship. Journal of Clinical Oncology: Official Journal of the American Society of Clinical Oncology. 2010;28(26):4022–8. DOI: 10.1200/JCO.2010.28.4257
  79. 79. Jung CH, Kim EM, Park JK, Hwang SG, Moon SK, Kim WJ, Um HD. Bmal1 suppresses cancer cell invasion by blocking the phosphoinositide 3-kinase-Akt-MMP-2 signaling pathway. Oncology Reports. 2013;29(6):2109–13. DOI: 10.3892/or.2013.2381
  80. 80. Markoulli M, Papas E, Cole N, Holden BA. The diurnal variation of matrix metalloproteinase-9 and its associated factors in human tears. Investigative Ophthalmology & Visual Science. 2012;53(3):1479–84. DOI: 10.1167/iovs.11-8365
  81. 81. Carafa V, Nebbioso A, Altucci L. Sirtuins and disease: the road ahead. Frontiers in Pharmacology. 2012;3:4. DOI: 10.3389/fphar.2012.00004
  82. 82. Guarente L. Franklin H. Epstein lecture: sirtuins, aging, and medicine. The New England Journal of Medicine. 2011;364(23):2235–44. DOI: 10.1056/NEJMra1100831
  83. 83. Chalkiadaki A, Guarente L. The multifaceted functions of sirtuins in cancer. Nature Reviews Cancer. 2015;15(10):608–24. DOI: 10.1038/nrc3985
  84. 84. Yang H, Bi Y, Xue L, Wang J, Lu Y, Zhang Z, Chen X, Chu Y, Yang R, Wang R, Liu G. Multifaceted modulation of SIRT1 in cancer and inflammation. Critical Reviews in Oncogenesis. 2015;20(1–2):49–64.
  85. 85. Fang Y, Nicholl MB. A dual role for sirtuin 1 in tumorigenesis. Current Pharmaceutical Design. 2014;20(15):2634–6.
  86. 86. Luo J, Nikolaev AY, Imai S, Chen D, Su F, Shiloh A, Guarente L, Gu W. Negative control of p53 by Sir2alpha promotes cell survival under stress. Cell. 2001;107(2):137–48.
  87. 87. Vaziri H, Dessain SK, Ng Eaton E, Imai SI, Frye RA, Pandita TK, Guarente L, Weinberg RA. hSIR2(SIRT1) functions as an NAD-dependent p53 deacetylase. Cell. 2001;107(2):149–59.
  88. 88. Dai JM, Wang ZY, Sun DC, Lin RX, Wang SQ. SIRT1 interacts with p73 and suppresses p73-dependent transcriptional activity. Journal of Cellular Physiology. 2007;210(1):161–6. DOI: 10.1002/jcp.20831
  89. 89. Pediconi N, Guerrieri F, Vossio S, Bruno T, Belloni L, Schinzari V, Scisciani C, Fanciulli M, Levrero M. hSirT1-dependent regulation of the PCAF-E2F1-p73 apoptotic pathway in response to DNA damage. Molecular and Cellular Biology. 2009;29(8):1989–98. DOI: 10.1128/MCB.00552-08
  90. 90. Zhang Q, Wang SY, Fleuriel C, Leprince D, Rocheleau JV, Piston DW, Goodman RH. Metabolic regulation of SIRT1 transcription via a HIC1:CtBP corepressor complex. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(3):829–33. DOI: 10.1073/pnas.0610590104
  91. 91. Wang C, Chen L, Hou X, Li Z, Kabra N, Ma Y, Nemoto S, Finkel T, Gu W, Cress WD, Chen J. Interactions between E2F1 and SirT1 regulate apoptotic response to DNA damage. Nature Cell Biology. 2006;8(9):1025–31. DOI: 10.1038/ncb1468
  92. 92. Wong S, Weber JD. Deacetylation of the retinoblastoma tumour suppressor protein by SIRT1. The Biochemical Journal. 2007;407(3):451–60. DOI: 10.1042/BJ20070151
  93. 93. Ikenoue T, Inoki K, Zhao B, Guan KL. PTEN acetylation modulates its interaction with PDZ domain. Cancer Research. 2008;68(17):6908–12. DOI: 10.1158/0008-5472.CAN-08-1107
  94. 94. Yuan H, Wang Z, Li L, Zhang H, Modi H, Horne D, Bhatia R, Chen W. Activation of stress response gene SIRT1 by BCR-ABL promotes leukemogenesis. Blood. 2012;119(8):1904–14. DOI: 10.1182/blood-2011-06-361691
  95. 95. Guo JY, Xia B, White E. Autophagy-mediated tumor promotion. Cell. 2013;155(6):1216–9. DOI: 10.1016/j.cell.2013.11.019
  96. 96. Wei Y, Zou Z, Becker N, Anderson M, Sumpter R, Xiao G, Kinch L, Koduru P, Christudass CS, Veltri RW, Grishin NV, Peyton M, Minna J, Bhagat G, Levine B. EGFR-mediated Beclin 1 phosphorylation in autophagy suppression, tumor progression, and tumor chemoresistance. Cell. 2013;154(6):1269–84. DOI: 10.1016/j.cell.2013.08.015
  97. 97. Galluzzi L, Pietrocola F, Levine B, Kroemer G. Metabolic control of autophagy. Cell. 2014;159(6):1263–76. DOI: 10.1016/j.cell.2014.11.006
  98. 98. Lee IH, Cao L, Mostoslavsky R, Lombard DB, Liu J, Bruns NE, Tsokos M, Alt FW, Finkel T. A role for the NAD-dependent deacetylase Sirt1 in the regulation of autophagy. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(9):3374–9. DOI: 10.1073/pnas.0712145105
  99. 99. Yuan J, Minter-Dykhouse K, Lou Z. A c-Myc-SIRT1 feedback loop regulates cell growth and transformation. The Journal of Cell Biology. 2009;185(2):203–11. DOI: 10.1083/jcb.200809167
  100. 100. Rajamohan SB, Pillai VB, Gupta M, Sundaresan NR, Birukov KG, Samant S, Hottiger MO, Gupta MP. SIRT1 promotes cell survival under stress by deacetylation-dependent deactivation of poly(ADP-ribose) polymerase 1. Molecular and Cellular Biology. 2009;29(15):4116–29. DOI: 10.1128/MCB.00121-09
  101. 101. Ming M, Shea CR, Guo X, Li X, Soltani K, Han W, He YY. Regulation of global genome nucleotide excision repair by SIRT1 through xeroderma pigmentosum C. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(52):22623–8. DOI: 10.1073/pnas.1010377108
  102. 102. Oberdoerffer P, Michan S, McVay M, Mostoslavsky R, Vann J, Park SK, Hartlerode A, Stegmuller J, Hafner A, Loerch P, Wright SM, Mills KD, Bonni A, Yankner BA, Scully R, Prolla TA, Alt FW, Sinclair DA. SIRT1 redistribution on chromatin promotes genomic stability but alters gene expression during aging. Cell. 2008;135(5):907–18. DOI: 10.1016/j.cell.2008.10.025
  103. 103. Zhu X, Liu Q, Wang M, Liang M, Yang X, Xu X, Zou H, Qiu J. Activation of Sirt1 by resveratrol inhibits TNF-alpha induced inflammation in fibroblasts. PLoS One. 2011;6(11):e27081. DOI: 10.1371/journal.pone.0027081
  104. 104. Zhang Z, Lowry SF, Guarente L, Haimovich B. Roles of SIRT1 in the acute and restorative phases following induction of inflammation. The Journal of Biological Chemistry. 2010;285(53):41391–401. DOI: 10.1074/jbc.M110.174482
  105. 105. Yoshizaki T, Schenk S, Imamura T, Babendure JL, Sonoda N, Bae EJ, Oh DY, Lu M, Milne JC, Westphal C, Bandyopadhyay G, Olefsky JM. SIRT1 inhibits inflammatory pathways in macrophages and modulates insulin sensitivity. American Journal of Physiology: Endocrinology and Metabolism. 2010;298(3):E419-28. DOI: 10.1152/ajpendo.00417.2009
  106. 106. Yeung F, Hoberg JE, Ramsey CS, Keller MD, Jones DR, Frye RA, Mayo MW. Modulation of NF-kappaB-dependent transcription and cell survival by the SIRT1 deacetylase. The EMBO Journal. 2004;23(12):2369–80. DOI: 10.1038/sj.emboj.7600244
  107. 107. Nie Y, Erion DM, Yuan Z, Dietrich M, Shulman GI, Horvath TL, Gao Q. STAT3 inhibition of gluconeogenesis is downregulated by SirT1. Nature Cell Biology. 2009;11(4):492–500. DOI: 10.1038/ncb1857
  108. 108. Zhang R, Chen HZ, Liu JJ, Jia YY, Zhang ZQ, Yang RF, Zhang Y, Xu J, Wei YS, Liu DP, Liang CC. SIRT1 suppresses activator protein-1 transcriptional activity and cyclooxygenase-2 expression in macrophages. The Journal of Biological Chemistry. 2010;285(10):7097–110. DOI: 10.1074/jbc.M109.038604
  109. 109. Zhang J, Lee SM, Shannon S, Gao B, Chen W, Chen A, Divekar R, McBurney MW, Braley-Mullen H, Zaghouani H, Fang D. The type III histone deacetylase Sirt1 is essential for maintenance of T cell tolerance in mice. The Journal of Clinical Investigation. 2009;119(10):3048–58. DOI: 10.1172/JCI38902
  110. 110. Beier UH, Wang L, Bhatti TR, Liu Y, Han R, Ge G, Hancock WW. Sirtuin-1 targeting promotes Foxp3+ T-regulatory cell function and prolongs allograft survival. Molecular and Cellular Biology. 2011;31(5):1022–9. DOI: 10.1128/MCB.01206-10
  111. 111. Chen X, Lu Y, Zhang Z, Wang J, Yang H, Liu G. Intercellular interplay between Sirt1 signalling and cell metabolism in immune cell biology. Immunology. 2015;145(4):455–67. DOI: 10.1111/imm.12473
  112. 112. Calvanese V, Lara E, Suarez-Alvarez B, Abu Dawud R, Vazquez-Chantada M, Martinez-Chantar ML, Embade N, Lopez-Nieva P, Horrillo A, Hmadcha A, Soria B, Piazzolla D, Herranz D, Serrano M, Mato JM, Andrews PW, Lopez-Larrea C, Esteller M, Fraga MF. Sirtuin 1 regulation of developmental genes during differentiation of stem cells. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(31):13736–41. DOI: 10.1073/pnas.1001399107
  113. 113. Li L, Osdal T, Ho Y, Chun S, McDonald T, Agarwal P, Lin A, Chu S, Qi J, Hsieh YT, Dos Santos C, Yuan H, Ha TQ, Popa M, Hovland R, Bruserud O, Gjertsen BT, Kuo YH, Chen W, Lain S, Mc Cormack E, Bhatia R. SIRT1 activation by a c-MYC oncogenic network promotes the maintenance and drug resistance of human FLT3-ITD acute myeloid leukemia stem cells. Cell Stem Cell. 2014;15(4):431–46. DOI: 10.1016/j.stem.2014.08.001
  114. 114. Lin L, Zheng X, Qiu C, Dongol S, Lv Q, Jiang J, Kong B, Wang C. SIRT1 promotes endometrial tumor growth by targeting SREBP1 and lipogenesis. Oncology Reports. 2014;32(6):2831–5. DOI: 10.3892/or.2014.3521
  115. 115. Lee JS, Park JR, Kwon OS, Lee TH, Nakano I, Miyoshi H, Chun KH, Park MJ, Lee HJ, Kim SU, Cha HJ. SIRT1 is required for oncogenic transformation of neural stem cells and for the survival of "cancer cells with neural stemness" in a p53-dependent manner. Neuro-oncology. 2015;17(1):95–106. DOI: 10.1093/neuonc/nou145
  116. 116. Mao B, Hu F, Cheng J, Wang P, Xu M, Yuan F, Meng S, Wang Y, Yuan Z, Bi W. SIRT1 regulates YAP2-mediated cell proliferation and chemoresistance in hepatocellular carcinoma. Oncogene. 2014;33(11):1468–74. DOI: 10.1038/onc.2013.88
  117. 117. Rando TA. Stem cells, ageing and the quest for immortality. Nature. 2006;441(7097):1080–6. DOI: 10.1038/nature04958
  118. 118. Calvanese V, Fraga MF. SirT1 brings stemness closer to cancer and aging. Aging. 2011;3(2):162–7.
  119. 119. Choi SE, Kemper JK. Regulation of SIRT1 by microRNAs. Molecules and Cells. 2013;36(5):385–92. DOI: 10.1007/s10059-013-0297-1
  120. 120. Kim EJ, Kho JH, Kang MR, Um SJ. Active regulator of SIRT1 cooperates with SIRT1 and facilitates suppression of p53 activity. Molecular Cell. 2007;28(2):277–90. DOI: 10.1016/j.molcel.2007.08.030
  121. 121. Kim JE, Chen J, Lou Z. DBC1 is a negative regulator of SIRT1. Nature. 2008;451(7178):583–6. DOI: 10.1038/nature06500
  122. 122. Yamakuchi M, Lowenstein CJ. MiR-34, SIRT1 and p53: the feedback loop. Cell Cycle. 2009;8(5):712–5.
  123. 123. Akao Y, Noguchi S, Iio A, Kojima K, Takagi T, Naoe T. Dysregulation of microRNA-34a expression causes drug-resistance to 5-FU in human colon cancer DLD-1 cells. Cancer Letters. 2011;300(2):197–204. DOI: 10.1016/j.canlet.2010.10.006
  124. 124. Amirkhah R, Farazmand A, Irfan-Maqsood M, Wolkenhauer O, Schmitz U. The role of microRNAs in the resistance to colorectal cancer treatments. Cell and Molecular Biology (Noisy-le-grand). 2015;61(6):17–23.
  125. 125. Tarantino C, Paolella G, Cozzuto L, Minopoli G, Pastore L, Parisi S, Russo T. miRNA 34a, 100, and 137 modulate differentiation of mouse embryonic stem cells. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology. 2010;24(9):3255–63. DOI: 10.1096/fj.09-152207
  126. 126. Aranha MM, Santos DM, Sola S, Steer CJ, Rodrigues CM. miR-34a regulates mouse neural stem cell differentiation. PLoS One. 2011;6(8):e21396. DOI: 10.1371/journal.pone.0021396
  127. 127. Lee YL, Peng Q, Fong SW, Chen AC, Lee KF, Ng EH, Nagy A, Yeung WS. Sirtuin 1 facilitates generation of induced pluripotent stem cells from mouse embryonic fibroblasts through the miR-34a and p53 pathways. PLoS One. 2012;7(9):e45633. DOI: 10.1371/journal.pone.0045633
  128. 128. Xu Z, Zhang L, Fei X, Yi X, Li W, Wang Q. The miR-29b-Sirt1 axis regulates self-renewal of mouse embryonic stem cells in response to reactive oxygen species. Cellular Signalling. 2014;26(7):1500–5. DOI: 10.1016/j.cellsig.2014.03.010
  129. 129. Nalls D, Tang SN, Rodova M, Srivastava RK, Shankar S. Targeting epigenetic regulation of miR-34a for treatment of pancreatic cancer by inhibition of pancreatic cancer stem cells. PLoS One. 2011;6(8):e24099. DOI: 10.1371/journal.pone.0024099
  130. 130. Duan K, Ge YC, Zhang XP, Wu SY, Feng JS, Chen SL, Zhang LI, Yuan ZH, Fu CH. miR-34a inhibits cell proliferation in prostate cancer by downregulation of SIRT1 expression. Oncology Letters. 2015;10(5):3223–7. DOI: 10.3892/ol.2015.3645
  131. 131. Ye Z, Fang J, Dai S, Wang Y, Fu Z, Feng W, Wei Q, Huang P. MicroRNA-34a induces a senescence-like change via the down-regulation of SIRT1 and up-regulation of p53 protein in human esophageal squamous cancer cells with a wild-type p53 gene background. Cancer Letters. 2016;370(2):216–21. DOI: 10.1016/j.canlet.2015.10.023
  132. 132. Ma W, Xiao GG, Mao J, Lu Y, Song B, Wang L, Fan S, Fan P, Hou Z, Li J, Yu X, Wang B, Wang H, Xu F, Li Y, Liu Q, Li L. Dysregulation of the miR-34a-SIRT1 axis inhibits breast cancer stemness. Oncotarget. 2015;6(12):10432–44. DOI: 10.18632/oncotarget.3394
  133. 133. Wang Z, Chen CC, Chen W. CD150 Side Population Defines Leukemia Stem Cells in a BALB/c Mouse Model of CML and Is Depleted by Genetic Loss of SIRT1. Stem Cells. 2015. DOI: 10.1002/stem.2218
  134. 134. Asher G, Gatfield D, Stratmann M, Reinke H, Dibner C, Kreppel F, Mostoslavsky R, Alt FW, Schibler U. SIRT1 regulates circadian clock gene expression through PER2 deacetylation. Cell. 2008;134(2):317–28. DOI: 10.1016/j.cell.2008.06.050
  135. 135. Nakahata Y, Kaluzova M, Grimaldi B, Sahar S, Hirayama J, Chen D, Guarente LP, Sassone-Corsi P. The NAD+−dependent deacetylase SIRT1 modulates CLOCK-mediated chromatin remodeling and circadian control. Cell. 2008;134(2):329–40. DOI: 10.1016/j.cell.2008.07.002
  136. 136. Chung S, Yao H, Caito S, Hwang JW, Arunachalam G, Rahman I. Regulation of SIRT1 in cellular functions: role of polyphenols. Archives of Biochemistry and Biophysics. 2010;501(1):79–90. DOI: 10.1016/j.abb.2010.05.003
  137. 137. Volonte D, Zou H, Bartholomew JN, Liu Z, Morel PA, Galbiati F. Oxidative stress-induced inhibition of Sirt1 by caveolin-1 promotes p53-dependent premature senescence and stimulates the secretion of interleukin 6 (IL-6). The Journal of Biological Chemistry. 2015;290(7):4202–14. DOI: 10.1074/jbc.M114.598268
  138. 138. Ranieri D, Avitabile D, Shiota M, Yokomizo A, Naito S, Bizzarri M, Torrisi MR. Nuclear redox imbalance affects circadian oscillation in HaCaT keratinocytes. The International Journal of Biochemistry & Cell Biology. 2015;65:113–24. DOI: 10.1016/j.biocel.2015.05.018
  139. 139. Villalba JM, Alcain FJ. Sirtuin activators and inhibitors. Biofactors. 2012;38(5):349–59. DOI: 10.1002/biof.1032
  140. 140. Kozako T, Suzuki T, Yoshimitsu M, Arima N, Honda S, Soeda S. Anticancer agents targeted to sirtuins. Molecules. 2014;19(12):20295–313. DOI: 10.3390/molecules191220295
  141. 141. Hsu YF, Sheu JR, Lin CH, Yang DS, Hsiao G, Ou G, Chiu PT, Huang YH, Kuo WH, Hsu MJ. Trichostatin A and sirtinol suppressed survivin expression through AMPK and p38MAPK in HT29 colon cancer cells. Biochimica et biophysica acta. 2012;1820(2):104–15. DOI: 10.1016/j.bbagen.2011.11.011
  142. 142. Rotili D, Tarantino D, Nebbioso A, Paolini C, Huidobro C, Lara E, Mellini P, Lenoci A, Pezzi R, Botta G, Lahtela-Kakkonen M, Poso A, Steinkuhler C, Gallinari P, De Maria R, Fraga M, Esteller M, Altucci L, Mai A. Discovery of salermide-related sirtuin inhibitors: binding mode studies and antiproliferative effects in cancer cells including cancer stem cells. Journal of Medicinal Chemistry. 2012;55(24):10937–47. DOI: 10.1021/jm3011614
  143. 143. Sonnemann J, Marx C, Becker S, Wittig S, Palani CD, Kramer OH, Beck JF. p53-dependent and p53-independent anticancer effects of different histone deacetylase inhibitors. British Journal of Cancer. 2014;110(3):656–67. DOI: 10.1038/bjc.2013.742
  144. 144. Lee SH, Kim MJ, Kim DW, Kang CD, Kim SH. Amurensin G enhances the susceptibility to tumor necrosis factor-related apoptosis-inducing ligand-mediated cytotoxicity of cancer stem-like cells of HCT-15 cells. Cancer Science. 2013;104(12):1632–9. DOI: 10.1111/cas.12299
  145. 145. Kumazaki M, Noguchi S, Yasui Y, Iwasaki J, Shinohara H, Yamada N, Akao Y. Anti-cancer effects of naturally occurring compounds through modulation of signal transduction and miRNA expression in human colon cancer cells. The Journal of Nutritional Biochemistry. 2013;24(11):1849–58. DOI: 10.1016/j.jnutbio.2013.04.006
  146. 146. Reiter RJ, Tan DX, Galano A. Melatonin: exceeding expectations. Physiology (Bethesda). 2014;29(5):325–33. DOI: 10.1152/physiol.00011.2014
  147. 147. Zamfir Chiru AA, Popescu CR, Gheorghe DC. Melatonin and cancer. Journal of Medicine and Life. 2014;7(3):373–4.
  148. 148. Jung-Hynes B, Schmit TL, Reagan-Shaw SR, Siddiqui IA, Mukhtar H, Ahmad N. Melatonin, a novel Sirt1 inhibitor, imparts antiproliferative effects against prostate cancer in vitro in culture and in vivo in TRAMP model. Journal of Pineal Research. 2011;50(2):140–9. DOI: 10.1111/j.1600-079X.2010.00823.x
  149. 149. Cheng Y, Cai L, Jiang P, Wang J, Gao C, Feng H, Wang C, Pan H, Yang Y. SIRT1 inhibition by melatonin exerts antitumor activity in human osteosarcoma cells. European Journal of Pharmacology. 2013;715(1–3):219–29. DOI: 10.1016/j.ejphar.2013.05.017
  150. 150. Kannen V, Marini T, Zanette DL, Frajacomo FT, Silva GE, Silva WAJr., Garcia SB. The melatonin action on stromal stem cells within pericryptal area in colon cancer model under constant light. Biochemical and Biophysical Research Communications. 2011;405(4):593–8. DOI: 10.1016/j.bbrc.2011.01.074
  151. 151. Zu G, Ji A, Zhou T, Che N. Clinicopathological significance of SIRT1 expression in colorectal cancer: A systematic review and meta analysis. The International Journal of Surgery. 2016;26:32–7. DOI: 10.1016/j.ijsu.2016.01.002
  152. 152. Vellinga TT, Borovski T, de Boer VC, Fatrai S, van Schelven S, Trumpi K, Verheem A, Snoeren N, Emmink BL, Koster J, Rinkes IH, Kranenburg O. SIRT1/PGC1alpha-dependent increase in oxidative phosphorylation supports chemotherapy resistance of colon cancer. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research. 2015;21(12):2870–9. DOI: 10.1158/1078-0432.CRC-14-2290

Written By

Sandra Ríos-Arrabal, José Antonio Muñoz-Gámez, Sergio Manuel Jiménez-Ruíz, Jorge Casado-Ruíz, Francisco Artacho-Cordón and Josefa León-López

Submitted: 28 September 2015 Reviewed: 29 January 2016 Published: 07 September 2016