ANTIOXIDANTS & REDOX SIGNALING Volume 00, Number 00, 2014 ª Mary Ann Liebert, Inc. DOI: 10.1089/ars.2014.5854

FORUM REVIEW ARTICLE

New Insights into the Connection Between Histone Deacetylases, Cell Metabolism, and Cancer Ferdinando Chiaradonna,1,2 Claudia Cirulli,1,2,* Roberta Palorini,1,2,* Giuseppina Votta,1,2,* and Lilia Alberghina1,2

Abstract

Significance: Histone deacetylases (HDACs) activity and cell metabolism are considered important targets for cancer therapy, as both are deregulated and associated with the onset and maintenance of tumors. Recent Advances: Besides the classical function of HDACs as HDAC enzymes controlling the transcription, it is becoming increasingly evident that these proteins are involved in the regulation of several other cellular processes by their ability to deacetylate hundreds of proteins with different functions in both the cytoplasm and the nucleus. Importantly, recent high-throughput studies have identified as important target proteins several enzymes involved in different metabolic pathways. Conversely, it has been also shown that metabolic intermediates may control HDACs activity. Consequently, the acetylation/deacetylation of metabolic enzymes and the ability of metabolic intermediates to modulate HDACs may represent a cross-talk connecting cell metabolism, transcription, and other HDACs-controlled processes in physiological and pathological conditions. Critical Issues: Since metabolic alterations and HDACs deregulation are important cancer hallmarks, disclosing connections among them may improve our understanding on cancer mechanisms and reveal novel therapeutic protocols against this disease. Future Directions: High-throughput metabolic studies performed by using more sophisticated technologies applied to the available models of conditional deletion of HDACs in cell lines or in mice will fill the gap in the current understanding and open directions for future research. Antioxid. Redox Signal. 00: 000–000.

deacetylation are histone deacetylases (HDACs). The balance of protein acetylation and deacetylation plays a critical role in the regulation of gene expression as well as in several other cellular processes. Indeed, dysfunctions of the HATs/ HDACs balance are associated with various diseases such as cancer (144), diabetes, asthma, cardiac hypertrophy, retroviral pathogenesis, and neurodegenerative disorders (145). The goal of this article is to review current understanding and to highlight the most recent advances with regard to HDACs and their connection with cell metabolism, especially in cancer. Among HDACs, the sirtuins family has long been recognized as a target and regulator of cell metabolism; in fact, several studies have extensively described the role of these proteins in controlling cellular metabolism (148); therefore, they will not be covered in this review. Excluding sirtuins,

Introduction

H

istones and a large number of proteins involved in a wide range of cellular processes can be post-translationally acetylated. This post-translational modification involves the transfer of an acetyl group from acetyl-coenzyme A (CoA), a metabolic intermediate, on the e-amino groups of specific lysine residues. Acetylation changes the electrostatic state of lysine from positive to neutral and increases the size of the amino-acid side chain, changing DNA–protein and protein–protein interaction, subcellular localization, transcriptional activity, stability, and involvement in signaling pathways of acetylated proteins (22, 88). The transfer of the acetyl group on specific lysines is catalyzed by a group of enzymes known as histone acetyltransferases (HATs) (23). The enzymes responsible for 1

SYSBIO Centre of Systems Biology, Milan, Italy. Department of Biotechnology and Biosciences, University of Milano-Bicocca, Milan, Italy. *These three authors contributed equally to this work.

2

1

2

less information has been generated regarding non-sirtuins or classical HDACs and cell metabolism. This is extremely important, as in recent years, several HDACs inhibitors are becoming a promising class of anti-cancer drug targets, already approved by FDA or in evaluation in clinical trials (169). In this regard, the exploration of the role of these proteins in controlling cell metabolism as well as whether metabolism controls HDACs assume an important meaning, especially in cancer cells, as the recent revival of interest in cell metabolism alterations as a key hallmark of tumorigenesis (84). This finding, indeed, has propelled the idea that metabolism could control several aspects of tumor biology and has led several researchers to the identification and the use in clinical trials of anti-cancer drugs targeting metabolism (110). Therefore, the knowledge on the connections between deacetylation and cell metabolism is a challenge to improve our understanding of cancer mechanisms and to develop novel therapeutic protocols against this disease. Classification, Expression, Regulation, and Function of HDACs

Mammalian genomes encode 11 classical HDACs (from here indicated only as HDACs) that are grouped into three classes according to their phylogenetic conservation, namely class I, II, and IV, characterized by a highly conserved deacetylase domain requiring a Zn2 + ion as cofactor (57) (Fig. 1). Class I

This class is composed by HDAC1, 2, 3, and 8. Their expression is ubiquitous, and they show a different subcel-

CHIARADONNA ET AL.

lular localization. HDAC1 and HDAC2, which have partially redundant roles, are localized exclusively in the nucleus (86, 121) (Fig. 1). HDAC3 and HDAC8 localize both in the cytoplasm and in the nucleus (158a, 171). HDACs do not bind DNA directly, but interact with DNA through multi-proteins complexes, including co-repressors and co-activators. HDAC1 and 2 are often associated together in multi-protein co-repressor complexes formed by suppressor of defective silencing 3 protein, nucleosome-remodeling deacetylase, and Co-repressor for element-1-silencing transcription factor (190). HDAC3 is able to form oligomers both in vitro and in vivo with other HDACs, itself, and nuclear receptor corepressor (N-CoR) protein (189). Less information is available for HDAC8, as recent reports demonstrated that recombinant purified HDAC8 catalyzes deacetylation and displays substrate selectivity also in the absence of additional protein cofactors (35, 183). Class I HDACs null mice are lethal. HDAC1 knockout (KO) die before embryonic day 10.5 for severe growth retardation (99, 121), and HDAC2 KO succumb during the perinatal period due to cardiac defects (121). HDAC3 KO die before E9.5 for defects in gastrulation, while the conditional deletion in the liver and heart disrupts lipid/cholesterol homeostasis and glycogen storage in the liver (91, 122). HDAC8 KO die within 4–6 h after birth due to deficiency of cranial neural crest cells (62) (Fig. 1). Class I members possess a simple structure consisting of the conserved deacetylase domain and short N- and Cterminus. As shown in Figure 2, HDAC1 and 2 possess the dimerization domain, which is necessary for their association and a nuclear localization signal (NLS); HDAC3 contains both nuclear export signal (NES) and NLS; and HDAC8

FIG. 1. Schematic representation of classical/zincdependent HDACs family. Cellular localization, possibility to form nuclear corepressor complexes, tissue expression, and effects of HDAC knockout in mice are reported for each classical/ zinc-dependent HDACs. B, brain; C, cytoplasm; E, embryonic day; H, heart; HDAC, histone deacetylase; K, kidney; L, liver; N, nucleus; ND, not determined; P, postnatal day; Pa, pancreas; Pl, placenta; S, spleen; SM, skeletal muscle; {, lethal.

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

3

FIG. 2. General structural features of human HDACs family. Classical HDACs family is divided into four classes, namely I, IIA, IIB, and IV. The modular organization of HDACs is highlighted. All members of this family show a highly conserved catalytic domain differently located in the four classes. Excluding HDAC10 and 11, all HDACs possess an NLS. Class II is characterized along with HDAC3 by an NES. Class I and II possess a dimerization domain or protein–protein interaction domains. For each human HDAC, the amino acids’ protein length and the known post-translational modifications are reported. NES, nuclear export signal; NLS, nuclear localization signal. To see this illustration in color, the reader is referred to the web version of this article at www.liebertpub.com/ars

contains only an NLS (57). Class I HDACs are regulated by post-translational modifications (Fig. 2). HADC1 activity is increased by Casein kinase 2 (CK2)-mediated phosphorylation (139) and SUMOylation (18), while it is decreased by ubiquitination (196), cysteine carbonylation (150), and acetylation (187). HDAC2 activity is increased by CK2phosphorylation (139) and decreased by s-nitrosylation (153), tyrosine-nitrosylation (129), and ubiquitination that specifically leads to its degradation (196). HDAC3 activity is stimulated by CK2-phosphorylation and antagonized by protein phosphatase 4-dephosphorylation (197). HDAC8 activity is decreased on protein kinase A (PKA)-phosphorylation (103). Class II

Class II members are further subdivided in class IIA and IIB (Fig. 1). HDACs 4, 5, 7, and 9, comprising class IIA, are able to shuttle between the cytoplasm and the nucleus; while HDACs 6 and 10 make up class IIB, mainly cytoplasmic. Class IIA. HDACs 4, 5, and 7 interact with silencing mediator of retinoid and thyroid receptors/N-CoR, Bcl-6interacting co-repressor, C-terminal-binding protein 1, and

HDAC3 in the nucleus (43). Class II HDACs expression is tissue specific; they are implicated in the development and differentiation of cardiac and skeletal striated muscle. HDAC4 KO die within the second week of life owing to ectopic ossification, while HDAC5 deletion causes cardiac defects, but mice are viable and fertile (17). HDAC7 KO results in embryonic lethality, owing to a loss of endothelial integrity (63). HDAC9 KO are viable but show severe cardiac defects (17) (Fig. 1). Class II members have the catalytic domain located at the C-terminus, NLS, NES (125) and a protein–protein interaction domain located at N-terminus, involved in the association with different molecular partners (Fig. 2). Class IIA localization is regulated by specific post-translational modifications. HDAC4 phosphorylation by calcium/calmodulindependent kinase (CaMK) and salt-inducible kinase 1 (SIK1) creates docking sites for nuclear export (59), while PKA triggers its proteolytic cleavage antagonizing the action of CaMK and generating a product that relocalizes into the nucleus (7). HDAC4 activity is reduced by SUMOylation (89). HDAC5 results in a strong nuclear accumulation after PKA-phosphorylation (16). HDAC7 may be phosphorylated by CaMK and protein kinase C (PKC) that impair nuclear import (68); the same effect is also yielded through HDAC7

4

CHIARADONNA ET AL.

proteolytic cleavage (149). Three different isoforms of HDAC9 generated by alternative splicing, with conserved structure but distinct cellular localization and cell specificity, have been found (136). HDAC9 SUMOylation reduces its activity (89), while phosphorylation by PKC kinase impairs the nuclear import (68). Class IIB. HDAC6 regulates cell motility, adhesion (198), and chaperone function (95). HDAC6-deficient mice are viable and fertile with hyperacetylated tubulin in most tissues (199) (Fig. 1). HDAC10, present in two variants, is the most recently discovered member of the class II. It interacts with different other HDACs, suggesting that it might exert cellular functions independently from its deacetylase activity by recruiting other deacetylases (42) (Fig. 1). HDAC6 contains two catalytic domains arranged in tandem and NLS and NES both located at N-terminus (Fig. 2). HDAC6 activity is increased on phosphorylation by CK2 (177) and oncogenic Aurora A kinase (142). HDAC5 and 6 have been also identified to be poly-ubiquitinated, but this modification does not target them for proteasomal degradation (130). HDAC10 has a catalytic domain on its N-terminus, and an NES and a putative second catalytic domain (Leucine-rich domain) on the C-terminus (Fig. 2). Class IV

Class IV comprises only HDAC11, which is localized in the nucleus and is not found to reside in any of the known HDAC complexes (49). It is structurally related to both class I and II and in particular to HDAC3 and HDAC8. HDAC11 contains a catalytic domain located at the N-terminus. Little information is available about its expression and function, but in recent years, some evidence of a role in the regulation of the immune system has been collected (167). HDACs targets Histone substrates. DNA is organized into chromatin. The basic subunit of chromatin is the nucleosome, which is composed of 147 bp of DNA, coiled around an octamer of histones, with two molecules each of histone H2A, H2B, H3, and H4. Histone H1 associates with chromatin outside the core octamer unit and regulates the higher order of the chromatin structure (195). Each histone in the nucleosome has lysine-rich flexible tails, and DNA accessibility is controlled by post-translational modifications of some of these lysines (Fig. 3). In particular, their deacetylation, resulting in an increase of the affinity between histones and DNA, makes gene transcription more difficult. Non-histone targets. Histones are not the only acetylated/deacetylated proteins. To date, more than 2000 acetylated proteins with various functions and involved in a wide range of cellular processes, including nuclear transport, chromatin remodeling, cytoskeleton rearrangements, cell cycle, DNA repair, and cell metabolism, have been identified (119) (Fig. 3). In particular, proteomic approaches have permitted the identification of a large number of acetylated metabolic proteins (Fig. 3) (70, 186), whose regulation is dependent by deacetylating HDACs activity. Conversely, a growing number of studies also report a fine regulation of HDACs by metabolism (Fig. 3). In this scenario, metabolism and

FIG. 3. Schematic representation of HDACs substrates and regulated processes. HDACs family members remove the acetyl groups from the e-N-acetyl lysine amino acid not only on a histone tail in the nucleosomes (upper part of the circle, histone targets), but also on the lysines of other proteins, the so-called non-histone targets (lower part of the circle), which have different functions and are involved in several physiological processes such as chromatin remodeling, nuclear transport, cytoskeleton rearrangements, cell cycle, DNA repair, and metabolism. Metabolism, in particular, is reviewed in this article. It is regulated by HDACs, but, in turn, may regulate HDAC family members.

HDACs appear to have different levels of communication: direct control of metabolic enzymes activity by HDACsmediated deacetylation, direct control of HDACs catalytic activity by metabolites, direct control of HDACs catalytic activity by metabolic responsive signaling, and lastly direct control of HDACs catalytic activity by level of free acetyl groups into the cells (179). In an attempt to collect the knowledge relative to these different levels of communications, next we will report some old and more recent findings about how metabolism regulates and is regulated by HDACs in physiological and pathological conditions. HDACs and Metabolism: Metabolites, Diet, and Metabolic Responsive Kinases as Control Mechanisms of HDAC Activity Metabolites and diet

Recent studies suggest that several metabolites, generated by different intracellular biochemical pathways, may directly regulate HDAC activity (168). In vitro and in vivo activity of recombinant HDAC1 and HDAC2 or immunoprecipitated HDAC1- and HDAC2-containing complexes increases on addition of NADPH and CoA-derivates. In particular, acetylCoA, acetoacetyl-CoA, succinyl-CoA, butyryl-CoA, isobutyryl-CoA, glutaryl-CoA, HMG-CoA, malonyl-CoA, methylmalonyl-CoA, crotonyl-CoA, and methylcrotonyl-CoA (molecules generated from degradation of glucose, or involved in the biosynthesis of fatty acids and sterols or in the metabolism of several amino acids and ketone bodies utilization) increase HDACs activity between 1.5 and 3-fold (Fig. 4). On the contrary, similar kinetic analyses indicate that the

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

long chain fatty acid derivative palmitoyl-CoA as well as the free CoA inhibit HDAC1 and HDAC2 activity (Fig. 4). Importantly the metabolite concentration used in these HDAC activity assays are almost comparable with their physiological intracellular levels (168). Further examples of metabolic regulation on HDACs activity are represented by sphingosine-1 phosphate, a bioactive lipid produced by nuclear sphingolipid metabolism, and l-carnitine, a short-chain nitrogen-containing carboxylic acid involved in fatty acid b-oxidation, that are able to bind the active site and inhibit HDACs activity both in vitro and in vivo (64, 73). Interestingly, it has also been shown, in different cancer cells, that pyruvate and lactate may act as specific inhibitors of HDACs. In particular, pyruvate shows an inhibitory effect comparable to butyrate, known histone deacetylase inhibitor (HDACi), on HDAC1 and HDAC3 (164) and lactate deregulates genes resembling significantly those affected by known HDACi such as butyrate and trichostatin A (TSA) (101). Although the lactate IC50 values for HDACs appear higher as compared with known HDACi, it is likely that lactate also may act as HDACi at a concentration similar to the physiological one as that observed; for instance, in skeletal muscle during prolonged exercise or in tumors which prevalently use glycolysis (146, 170). Recently, it has been shown that low intracellular pH favors HDACs activity and a significant decrease of global histones acetylation. This effect, leading to the release of free acetate anions, prevents further pH reduction, as they are transported along with the acid H + protons to the extracellular space. Notably, internal pH is significantly influenced by a reduction of extracellular pH, as for example, on accumulation of lactic acid in cancer cells, suggesting an association between acidic cancer cell environment and low levels of histone acetylation. The mechanism by which pH influences the HDACs activity is still to be determined (114). As mentioned, butyrate, a short chain fatty acid that is formed in the human colon by bacterial fermentation of carbohydrates, has a non-competitive inhibitory effect on HDACs (25). Recent data have shown that butyrate at a concentration of 0.5 mM inhibits the growth of HCT116 co-

FIG. 4. Overview of the metabolites and the physical parameters affecting HDACs activity. In the presence of a water molecule, HDACs catalyzes the removal of the acetyl group from lysine, regenerating the e-amino group and releasing an acetate molecule. This reaction is modulated by reported metabolites or physical parameters through their activator or inhibitor function on HDACs.

5

lon cancer cells, while it has a stimulatory effect on the growth of non-cancerous cells, which is, however, lost at higher concentrations (2–5 mM) (34). The authors demonstrate that this effect is caused by the differences between normal and cancer cells in terms of energy metabolism. Normal cells, in fact, being characterized by an oxidative metabolism, also use a low dose of butyrate through the tricarboxylic acid (TCA) cycle generating acetyl-CoA that ultimately favors HAT activity and proliferation. Conversely, cancer colonocytes being characterized by a glycolytic metabolism use a relatively small amount of butyrate through TCA cycle, enabling it to accumulate inside of the nucleus where it acts as an HDAC inhibitor. At higher concentrations, butyrate accumulates into the nucleus and also acts as an inhibitor in normal cells (34, 81). b-Hydroxybutyrate (bOHB), the major source of energy for mammals, present in organisms at millimolar concentrations during prolonged exercise or starvation (141), has been also identified as a new physiological inhibitor of HDACs activity (154), in particular of HDAC1, HDAC3, and HDAC4 isoforms (154). An association between HDACs activity and metabolism has also been observed at the organismal level. In fact, metabolic diseases such as diabetes and obesity involve epigenetic modifications influenced by diets (137). Foods such as high-fat and simple sugars affect HDAC activity in different animal models. Japanese macaques, fed with a highfat diet during pregnancy, give birth to offspring with an increased histone H3 acetylation and decreased HDAC1 expression in the liver compared with macaques fed with a low-fat diet (1). The liver is a center of energy metabolism and the principal site of detoxification within the body; therefore, detrimental effects induced by an unhealthy diet can usually be observed in this organ for failure of correct tissue regeneration on partial hepatectomy (116). Mice and rats fed with a high-fat diet have shown impaired ability to regenerate liver on partial hepatectomy as compared with animals fed with a normal diet (29, 161). Notably, liver regenerative ability of high-fat diet animals or old mice hepatocyte proliferation are inversely correlated with hepatic HDAC1 protein levels (161, 173). Similarly, it has been

6

shown that hepatic HDAC3 higher expression levels are associated to metabolic syndrome after a high-fat diet (157). Similar phenotypes are obtained when mice are fed with a simple sugar diet, and this effect could be linked to HDACs regulation (124). Taken together, all these studies suggest that HDACs are regulated by intermediates of cell metabolism, are responsive to metabolic changes, and are directly involved in the epigenetic modifications induced by diet. Metabolic responsive kinases

Several findings demonstrate that not only the activity of HDACs against their substrates are physiologically relevant but also their function as substrates for other modifying enzymes, mainly kinases, is important for the development and maintenance of cellular homeostasis and cellular metabolism. In recent years, several studies have highlighted the role of metabolism in acetylation modifications through the regulation of HDACs activity. In this regard, some kinases, previously described for their ability to regulate HDACs activity through post-translational modifications, are known to be sensitive to exogenous and endogenous stimuli, leading to a change of intracellular metabolism (39, 200) (Fig. 5).

CHIARADONNA ET AL.

Adenosine monophosphate-activated protein kinase (AMPK) is a key metabolic kinase that is activated in response to a decline of cell energetic status and AMP/ATP ratio increase, which regulates both catabolic and anabolic processes. In primary hepatocytes and hepatoma cell lines, class IIa HDACs (HDAC4,5,7) are phosphorylated and excluded from the nucleus by AMPK family kinases (Fig. 5A), leading to the inhibition of gluconeogenic genes transcription (118). On the other hand, PKA kinase, which is involved in several cellular functions including regulation of glycogen, sugar, and lipid metabolism (78, 163), may inhibit the AMPK activity, permitting the accumulation of HDAC4,5,7 in the nucleus, where they act as scaffolds for HDAC3. The stimulation of HDAC3 leads to deacetylation and consequent activation of the transcription factor Forkhead box class O 1 (FoxO1) that induce the expression of gluconeogenic genes (118) (Fig. 5A). The inhibition of HDAC4 and HDAC5 by CaMKII and SIK1, a kinase that belongs to the AMPK family (77), leads to the activation of the glucose transporter GLUT4 and of peroxisome proliferator-activated receptor-c coactivator (PGC-1a) transcription (115, 127) and to the inhibition of glyconeogenesis (159), therefore controlling glucose metabolism and mitochondrial biogenesis (Fig. 5B).

FIG. 5. Overview of kinases involved in metabolic pathways that on metabolic stimuli regulate HDACs activity by phosphorylation. (A) In the fed state, liver kinases of AMPK family phosphorylate class IIA HDACs, inducing their interaction with 14-3-3 binding proteins. Such an interaction, leading to HDACs 4,5,7 cytoplasmic localization, induces their inactivation. In this condition, the transcription factor FoxO remains acetylated and inactive, therefore unable to induce gluconeogenesis gene transcription. Under fasting conditions, the hormone glucagon triggers the activation of PKA and the nuclear recruitment of class IIA HDACs, which, in turn, recruit the HDAC3 in order to deactylate and activate FoxO transcription factor, enabling the expression of gluconeogenic genes. (B) The energetic states and cytosolic Ca2 + levels regulate SIK1 and CaMK, respectively. On activation, these kinases phosphorylate class IIA HDACs, leading to 14-3-3 binding and inducing nuclear export and transcriptional repression of target genes. (C) Under hypoxic condition, CK2 phosphorylates HDAC1,2, promoting their inhibitory activity on VHL promoter and, hence, to the increased expression and stabilization of HIF1a protein. AMPK, adenosine monophosphate-activated protein kinase; CK2, casein kinase 2; CaMK, calcium/calmodulin-dependent kinase; FoxO, forkhead box class O; HIF1a, hypoxia-inducible factor 1a; PKA, protein kinase A; SIK1, salt-inducible kinase 1; VHL, E3-ubiquitin ligase von Hippel-Lindau.

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

In Hela cells subjected to hypoxia (Fig. 5C), HDAC1 and HDAC2 activity is stimulated by CK2-phosphorylation (139). HDACs activation leads to transcriptional repression of the E3-ubiquitin ligase von Hippel-Lindau (VHL) and, hence, to the stabilization of its main target, the transcriptional factor Hypoxia-inducible factor 1a (HIF1a) (87). Whether other kinases as well as other stimuli are able to control cell metabolism through the regulation of HDACs activity remains to be investigated; however, the ability of metabolic responsive kinases to regulate HDACs activity is a newly discovered important connection linking different metabolic states and transcriptional programs. HDACs and Cancer Metabolism Protein hypoacetylation and deregulated metabolism are distinctive features of tumors

Due to their important role in the regulation of gene transcription and protein activity, HDACs are essential for proper growth and behavior of cells and tissues. Therefore, deregulated function of HDACs can promote several developmental defects, as described in Figure 1, and also cancer onset and progression (53). Several studies have observed that tumors are characterized by the unbalance between protein acetylation and deacetylation (119). Although some tumors are characterized by a relative increase of acetylation, as a consequence of functional mutations and overexpression of HAT proteins (24, 133), several evidence indicates mainly an unbalance favoring the activity of HDACs, as a result of their over-expression (182) and/or aberrant recruitment to target promoters (50, 185), that leads to histone hypoacetylation. In this regard, significant evidence of the involvement of HDACs in cancer progression has been well documented through selective knockdown of single HDACs or by using specific HDAC inhibitors (63, 144). Interestingly, such approaches have also permitted to disclose specific targets and molecular mechanisms that link HDACs to tumorigenesis. As described earlier, HDACs have histone and non-histone targets and among them, several metabolic proteins or proteins involved in the regulation of the cellular metabolism have been identified. Indeed, recently, it has been shown that lysine acetylation is a prevalent modification in enzymes that catalyze intermediary metabolism (106). Importantly, such a modification, directed by an HAT and HDAC pair, appear to be synchronized according to growth conditions and regulate the metabolic enzyme activity and the metabolic fluxes (175). Therefore, deregulated HDACs could be involved in metabolic reprogramming that characterizes cancer cells as compared with normal ones (117). In this regard we would like to briefly summarize the most relevant among these metabolic differences (Fig. 6). Cancer cells are very often characterized by an intense glycolytic activity that leads to a large production of lactic acid even at normal oxygen availability, the so-called Warburg effect. Several reports have shown that this enhanced glycolysis is associated to an increase in the activity of the glycolytic enzymes such as hexokinase1 and 2 (HK1-2), pyruvate kinase M2 (PKM2), lactate dehydrogenase A (LDHA), and monocarboxylate transporter 4 or to the upregulation of glucose transporters GLUT1-3 (20). Warburg effect has also been related to the alteration of different oncogenic pathways such as protein kinase B (Akt or PKB) and

7

phosphoinositide 3-kinase, whose activation is sufficient to functionally drive glucose uptake and aerobic glycolysis (14, 31, 37). Several studies have also demonstrated that the enhanced glycolytic activity is a tumor cell metabolic strategy to survive in hypoxia by the induction of HIF1a transcription factor (13). Conversely, the switch to glycolysis of cancer cells is also promoted by the inactivation of tumor suppressor proteins. In fact, phosphatase and tensin homolog deleted in chromosome 10 (PTEN) and p53 inactivation, leading to chronic activation of Akt or to a reduction of mitochondrial respiration (9), respectively, favor the increase of glycolysis observed in cancer cells. Another metabolic adaptation of cancer cells is given by their tendency to increase glutamine consumption (30, 181); in fact, they often show an increased expression of membrane glutamine transporters as well as enzymes involved in the anabolic use of glutamine backbone such as glutaminase, glutamate dehydrogenase, aspartate amino transferase, and glutathione sinthetase (181). In addition, cancer cells often present an increased lipogenesis as reflected in the overexpression and hyperactivity of lipogenic enzymes such as ATP-citrate lyase, acetyl-coenzyme A carboxylase (ACC), or fatty acid synthase (97). Further metabolic differences between normal and cancer cells have been found in the mitochondrial respiratory system. In fact, cancer cell proliferation as well as tumor aggressiveness correlates with a lowmitochondrial respiratory chain activity (4, 19), established by the fact that the enhancement of oxidative phosphorylation (OXPHOS) activity reduces tumor growth (72, 147). Histones and non-histones deacetylation affects cellular metabolism through gene expression and protein function regulation

In a previous section, we already mentioned that HDACs might control metabolic genes through their effect on transcriptional machinery such as the activation of gluconeogenic and GLUT4 genes. Here, we would like to underline the regulatory activity of HDACs on the level of expression and/or activity of important proteins involved in cell metabolism control. Peroxisome proliferator-activated receptor-c coactivator.

PGC-1a is a transcription co-activator that plays a key role in the regulation of cellular energy metabolism. PGC-1a, in fact, in order to increase respiratory capacity of mitochondria, may stimulate mitochondrial biogenesis and/or coordinate organelle remodeling through mammalian target of rapamycin (mTOR) pathway (6, 10). Moreover, it promotes gluconeogenesis, participates in the regulation of both carbohydrate and lipid metabolism, and promotes the remodeling of muscle tissue to a fiber-type composition that is metabolically more oxidative and less glycolytic (143). PGC1a promoter is down-regulated in response to a change of cell energy demands by specific deacetylation of histone H3 HDAC5-mediated that leads to a reduced mitochondrial activity (27) (Fig. 6). We might suggest that the effects of HDACs deacetylation on PGC-1a could be considered an important mechanism to control cell bioenergetics that need further investigation. Phosphatase and tensin homolog deleted in chromosome 10. PTEN is a phosphatase that determines the inhibition of

8

CHIARADONNA ET AL.

FIG. 6. Metabolic reprogramming in cancer cells and role of HDACs in regulating metabolism. In the cartoon, the tumor metabolic features are represented. Red arrows indicate the increased flux throughout the reactions, while green arrows indicate the decreased flux. As shown and discussed in the text, many cancer cells up-regulate glycolysis (violet box), pentose phosphate pathway (green box), lipogenesis (burgundy box), and glutamine metabolism (yellow box) in order to sustain the anabolic needs for their high proliferative rate. On the contrary, many cancer cells down-regulate the mitochondrial oxidative metabolism, decreasing the conversion of pyruvate to acetyl-CoA, the flux throughout the TCA cycle (orange box), and OXPHOS (light blue box). In the red rectangles, the main enzymes often found up-regulated in cancer cells are indicated. Several factors (oncogenes and tumor suppressors) have been identified as central regulators of cancer metabolism. The principal ones, discussed in the text, are represented in the scheme; blue symbols represent the normal regulation (/ activation; x inhibition) of the metabolic pathways by the depicted factors. These factors regulate the expression or the activity of specific proteins but for simplicity in the scheme, we show their influence on the entire pathways. Near the factors, the colored squares indicate their regulation by the HDACs, as described throughout the text. Red squares indicate increased activity (caused by various mechanisms); green squares indicate decreased activity; and a gray square is used when the regulation has not yet been determined. A yellow square, specific for YY1, indicates that HDACs favor its repressor activity instead of the activator function. The numbers over the squares indicate the HDACs involved in the regulation (for example, 1 represents HDAC1); nd, not determined; OXPHOS, oxidative phosphorylation; TCA, tricarboxylic acid; YY1, Yin Yang 1. To see this illustration in color, the reader is referred to the web version of this article at www.liebertpub.com/ars the Akt signaling pathway. It is a tumor suppressor that is frequently mutated, deleted, or transcriptionally repressed in a large number of tumors (155). Several reports have indicated that an important mechanism for PTEN repression is represented by the HDACs-dependent deacetylation of its promoter (47, 48, 126, 132). In particular, such a repression is due to the recruitment of HDAC1 to its promoter by Sp1. This mechanism has been proved both in normal and in cancer cells and convincingly confirmed in experiments in which the treatment with TSA strongly increases PTEN expression (94, 128, 132). On the other hand, acetylation/deacetylation of PTEN protein has also been recognized as a mechanism to control its activity; in fact, it is positively regulated by HATs and by TSA treatment. However, the HDACs involved in the deacetylation of PTEN protein are not known, at least for the E1 A-associated protein p300 (p300)/CBP-associated factor (PCAF)-acetylated residues. Regardless of the mechanism, PTEN repression leads to the activation of the oncogenic Akt pathway and, thus, to a metabolic reprogramming (52) (Fig. 6). Unfortunately, no papers have yet directly addressed the

role of the axis HDACs-PTEN in the occurrence of cancer cell metabolic alterations. Hypoxia-inducible factor 1a. HIF1 is a transcriptional factor that regulates the expression of genes which are critical to cellular response and adaptation to hypoxia, among which there are several glycolytic enzymes, glucose transporters, and TCA cycle inhibitors such as pyruvate dehydrogenase kinase 1 (PDK1) (152), favoring glycolysis and glucose fermentation over mitochondrial OXPHOS. HIF1a is constantly synthesized but rapidly degraded under non-hypoxic conditions by the VHL/proteasome-mediated degradation (151). However, HIF1a is also regulated by acetylation/deacetylation. HDACs of class II are important for stability of HIF1a. It has been shown that HDAC7 co-translocates to the nucleus with HIF1a under hypoxic conditions, increasing its transcriptional activity (85). Moreover it has been shown that HDAC6 stabilizes HIF1a by deacetylation of the chaperone Hsp90. In accordance, HDAC6 inhibition results in hyperacetylation of Hsp90 that leads to HIF1a proteasomal

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

degradation (92). Conversely, HDAC4 deacetylates Nterminal lysine residues of HIF1a, increasing protein stability under hypoxia. In fact, HDAC4 knockdown increases HIF1a acetylation and attenuates cancer cell response to hypoxia, inhibiting the HIF1a-mediated transcriptional activation of specific genes, among which are the glycolytic genes (GLUT1 and LDHA), and significantly reducing the hypoxia-dependent lactic acid production (51). Therefore, HDACs have an essential role in the glycolytic phenotype induced by hypoxia especially in cancer cells, which are often under reduced oxygen availability (134). HIF1a is also expressed in tumors independently from the oxygen environment (20). In this regard, it has been observed that the Metastasis-associated protein 1, whose elevated expression is correlated with tumors aggressiveness (104), enhances the stability of HIF1a by recruiting HDAC1 and permitting the HIF1a deacetylation (193). Such a mechanism, in a context of HDACs over-expression characteristic of a large number of tumors (182), may also lead to HIF1a stabilization and relative metabolic effects in an oxygen-independent manner (Fig. 6). p53. p53 is the most commonly mutated tumor suppressor in human cancer. It codifies for a transcription factor that activates a large number of target genes which are involved in different processes, among which is cellular metabolism. In fact, p53, through its controlling role in specific proteins, for example, TP53-induced glycolysis and apoptosis regulator (9), cytochrome c oxidase 2 (113), and malic enzyme (ME1–2) (79), may regulate processes such as glycolysis (9), OXPHOS (113), pentose phosphate pathway (80), lipogenesis (79), NADPH production (79), glutamine metabolism (79, 158), and reactive oxygen species (ROS) levels (158). Therefore, its inactivation leads to a strong metabolic reprogramming that favors tumorigenesis (Fig. 6). It has been demonstrated that p53 activity is increased by acetylation at specific lysine residues in its C-terminal domain (60). While acetylation is essential for p53 function, class I HDACs deacetylation is necessary to maintain the protein in a functionally inactive state (83, 108). Acetylation/ deacetylation of p53 also regulates its stability. In fact, since the acetylated lysine residues in C-terminal are also the sites for the mouse double minute 2 homolog (MDM2)-mediated ubiquitination and degradation of p53, on acetylation such a process is prevented. Interestingly, MDM2, in order to induce p53 degradation, first recruits an HDAC1-containing complex that promotes p53 deacetylation (75). More recently, it has been observed that HDAC2 also modulates the p53 transcriptional activity, indeed its knockdown increases p53DNA binding activity through the modulation of histone acetylation and chromatin composition (66). Altogether, these observations strongly support a role of HDACs in the p53 loss of function that favors the cancer metabolic switch (Fig. 6). Cellular myelocytomatosis viral oncogene homolog. Cellular myelocytomatosis viral oncogene homolog

(c-Myc) is an oncoprotein commonly up-regulated in cancer (112). c-Myc has an essential role in regulating cancer cell metabolism. In particular, c-Myc favors the oxidation of glucose to lactate, increasing the expression of glucose transporters and glycolytic enzymes, suppressing mitochondrial respiration, and inducing PDK1 (28). Moreover, c-Myc

9

can stimulate glutaminolysis, conferring glutamine-addiction to the cells (180) (Fig. 6). c-Myc activity is strongly regulated through different mechanisms (12). Several reports have indicated the c-Myc-recruitment of different HATs or of HDAC1 and HDAC3 as the primary mechanism for the regulation of its ability to activate or repress several target genes (38, 44, 98, 166). c-Myc is also regulated by acetylation. In fact, it has been shown that its stability is reduced by acetylation (38). However, there are no direct studies addressing the role and the HDACs involved in the deacetylation of c-Myc. On the other hand, several evidence indicates that the HDACi may down-regulate c-Myc expression (71, 96). Since there is an important role of c-Myc in different aspects of cancer metabolism, the characterization of the role of HDACs in c-Myc activity could be a vast area of investigation, leading to new elements for understanding tumorigenesis and cancer metabolism. Signal transducer and activator of transcription 3. Signal transducer and activator of transcription 3 (STAT3) is a transcriptional factor that is activated by cytokines, growth factors, and oncogenes. On stimulation, phosphorylated STAT3 dimerizes and translocates into the nucleus or mitochondria to activate target genes (160). STAT3 controls different cellular processes, among which is cell metabolism (32, 55, 178) (Fig. 6). In fact, when phosphorylated at tyrosine 705, it translocates into the nucleus, where, through the transcriptional induction of HIF1a, it promotes aerobic glycolysis. In addition, also acting as a transcriptional repressor for the nuclear-encoded genes involved in mitochondrial function, it strongly reduces mitochondrial respiration (32). On the other hand, on phosphorylation at serine 727, STAT3 migrates into the mitochondria where it is able to modulate OXPHOS (55, 178), an effect that in Ras-transformed cells has been shown to favor aerobic glycolysis associated with transformation (55). Importantly, it has been demonstrated that STAT3 is also regulated by (de)acetylation. In particular, in different solid cancer cells, acetylation of lysine 685 is critical for STAT3 dimerization, DNA binding, and transcriptional activity, as HDAC1-3 activation has negative effects on the protein function (194). On the contrary, in diffuse large B-cell lymphoma, the activity of HDAC3 is necessary for the nuclear function of STAT3. In fact, inhibition or knockdown of HDAC3 increases STAT3 lysine 685 acetylation, preventing tyrosine 705 phosphorylation and STAT3 nuclear translocation (61). Therefore, the role of acetylation in regulating STAT3 appears important for its function (Fig. 6). No studies have been addressed so far to clarify the role of HDACs in the dual metabolic function of STAT3 into the nucleus or mitochondria. Yin Yang 1. Yin Yang 1 (YY1) is a sequence-specific DNA-binding transcription factor that is able to activate or repress gene expression depending on the receiving stimuli and the factor to which it is associated (54). Its transcriptional activator function depends on its association with p300, while the transcriptional repressor activity is mediated by HDAC1, 2, and 5 (156, 191). Growing evidence reports YY1 as a downstream effector of the mTOR pathway (74, 100). A genomic analysis reveals that mTOR balances the cell energy metabolism by modulating the ability of YY1 to interact with transcriptional cofactors, such as PGC-1a, that are involved

10

in energy metabolism regulation (6). In fact, knockdown of YY1 causes a significant decrease in mitochondrial gene expression, a decrease of mitochondrial respiration, and the lost of the mTOR-dependent gene activation (10, 26) (Fig. 6). Moreover, it has been shown that YY1 is a negative regulator of p53 protein, and a positive regulator of HIF1a given that its inhibition reduces the accumulation and the activity of HIF1a under hypoxic condition (184). Such an observation suggests the involvement of HDAC-induced repressor activity of YY1 in cancer metabolic switch. Forkhead box class Os. FoxO belongs to the Forkhead box family of transcription regulators of cell cycle, apoptosis, and response to oxidative stress. FoxO subclass comprises four members in mammals: FoxO1, FoxO3a, FoxO4, and FoxO6. FoxOs promote tumor suppression (131) and have relevant roles in metabolic pathways regulating ROS detoxification (93), energy metabolism, glucose homeostasis (58), and expression of nuclear mitochondrial encoding genes (41). FoxOs are acetylated. Acetylation results in a reduced DNA binding and in an alteration of their cellular localization and, in consequence, FoxOs are inactivated (165). Conversely, deacetylation, due to class I/IIA HDACs, results in a strong activation of the transcriptional activity of FoxOs proteins (118, 172). Mihaylova et al. (118) directly correlate FoxO1 and 3a deacetylation/activation by HDACs to glucose homeostasis in mice. Furthermore, Shimazu et al. (154) show that bOHB acts as an endogenous HDAC inhibitor on HDAC1, increasing FoxO3a activity and leading to protection from oxidative stress (Fig. 6). Adenosine monophosphate-activated protein kinase. As previously described, AMPK can regulate localization and activity of HDACs (115, 159). Recently, it has been shown that AMPK is regulated by p300 and HDAC1 (107). p300 acetylates and inhibits AMPK, while the deacetylation of the kinase by HDAC1 favors its interaction with the upstream liver kinase B1 that on phoshorylation favors the AMPK activation. Deacetylated activated AMPK can phosphorylate its downstream targets, among which is the enzyme ACC. In cancer cells, where the equilibrium between HATs and HDACs is often lost, the metabolic kinase AMPK could be, consequently, deregulated (Fig. 6). To date, the role of AMPK as tumor suppressor or oncogene is still controversial and not yet completely clarified, as it has been suggested that its role may change depending on the stage of the tumor (65). In fact, AMPK, in general, switches off the anabolic pathways (65) and has been recognized as a negative regulator of the Warburg effect (40) but, on the other hand, it appears, especially in established tumors, essential for cancer development because of its fundamental role in maintaining energy homeostasis and protecting from metabolic stress (65, 105). Therefore, it may be very interesting to further investigate its regulation by HDACs, also considering the increasing use of HDACi as well as AMPK activators (200) as anticancer drugs. Pyruvate kinase M2. PK regulates the final step of glycolysis. Alternate splicing leads to the generation of PKM2, which is often over-expressed in cancer cells, where it plays

CHIARADONNA ET AL.

a central role in metabolic reprogramming, in the regulation of gene expression, and in the subsequent cell cycle progression (188) (Fig. 6). Lv et al. (109) demonstrate that PKM2 is acetylated by PCAF under high glucose concentrations and that such a modification leads to its reduced activity and to its lysosomal-dependent degradation. This mechanism permits tumor cells to accumulate or redirect some glycolytic intermediates to other anabolic pathways, for example, serine synthesis (192), in order to sustain cell growth especially under a rich glucose condition. Moreover, the authors show that PKM2 acetylation is enhanced after treatment with the HDACi TSA suggesting an involvement of different HDACs families in PKM2 deacetylation. To date, no more information is reported on this point, so it is possible to speculate that PKM2 deacetylation and allosteric activation by fructose 1, 6-bisphosphate in cancer positively regulate PKM2 activity, and, as a consequence, favor fermentative glycolysis and, hence, the Warburg effect. Drugs inhibiting HDACs activity induce large metabolic rearrangement in cancer cells

HDACs have become one of the emerging targets for cancer therapy, and HDACi show promising anticancer activities through their ability to induce differentiation and/or apoptosis in cancer cell lines (11). However, the overall impact of these inhibitors widely varies depending on the cell type, the dose, and the targeting of factors outside of transcription. In this regard, in recent years, significant attention has been focused on the ability of HDACi to alter cell metabolism along with their effect on histone acetylation. In particular, it has been shown that valproic acid and suberoylanilide hydroxamic acid (SAHA) treatments of multiple myeloma cells disturb acetyl-CoA intracellular equilibrium, as a consequence of a decreased fatty acid b-oxidation and a parallel reduction of glucose uptake and utilization. These effects are the results of the ability of the two inhibitors to down-regulate the expression of GLUT1 and to inhibit the enzymatic activity of HK1 (176). In addition, an increased oxidative metabolism of amino acids has been observed, indicating that the inhibitors lead to a large metabolic reprogramming in cancer cells. Of note, these effects have been also observed in NB4 acute promyelocytic leukemia cells, suggesting a more general mechanism at least in non-solid tumors. Alteration of glucose metabolism following SAHA and TSA treatments has also been observed in different colon cancer cells lines, where HDACi induce a transcriptionaldependent reduction of epidermal growth factor receptor (EGFR) expression. Since EGFR has been reported to be involved in the stabilization of the active Sodium-glucose transport protein 1, such a reduction is also associated to a significant reduction of glucose uptake and to an increase of the HDACi-induced cell death. It should be noted that cell death is almost rescued by increasing extracellular glucose availability (21), suggesting that glucose metabolism alterations participate in the cell death effect of HDACi. A more detailed bioenergetics analysis, performed after treatment with sodium butyrate or TSA in lung cancer cells, has further confirmed the role of HDACs in affecting cancer cell metabolism. In fact, both inhibitors increase mitochondrial respiration over glycolysis. Such a change in cell metabolism is associated with a decreased and increased expression of

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

GLUT1 and GLUT3 respectively, an enhancement of mitochondrial HK, a stimulation of glucose-6-phosphate dehydrogenase activity, and an increase of oxygen consumption coupled to ATP production. Importantly, the authors suggest that the negative effect of HDACi on HIF1a expression could justify all the observations on these lung cancer cells (5). An association between the increase of histone acetylation, inhibition of cell cycle, and remodeling of cell metabolism has been also observed in colon (HT29) and prostate (PC3) cancer cells as well as in HT29 xenograft tumors on Belinostat treatment, an HDACi similar to SAHA. This inhibitor induces an increase in the phosphocoline and in branchedchain amino acids (comprising valine, leucine, and isoleucine) levels and a reduction of glucose flux to lactate in favor of a rise in alanine formation (8). Altogether, these reports strongly support the role of HDACs in the positive control of cancer cell metabolism either through transcriptional regulation or through non-histone protein deacetylation. On the other hand, such a role has also been confirmed by using HDACi in a physiological condition. In fact, it has been shown that class I HDACi enhances whole-body energy expenditure, improves insulin sensitivity, and stimulates OXPHOS and mitochondrial function in mice skeletal muscle and adipose tissue. In contrast, class II HDACi do not exhibit these actions, suggesting that HDAC isoforms have specific energy metabolic targets in vivo (46, 176). Cancer cell metabolic reprogramming correlates with HDACs increased activity

As illustrated earlier, several metabolic intermediates are able to regulate HDACs activity both positively and negatively. Interestingly, several of these metabolites as well as physical changes are well correlated with the metabolic alterations observed in cancer cells (Fig. 7). Among the

FIG. 7. Cancer metabolism influences HDACs activity. In the scheme, some metabolic pathways as regulated in cancer cells are represented, indicating in red the up-regulated fluxes and in green the down-regulated ones. In the boxes, the physical parameters and the metabolic intermediates that have been shown to influence HDACs activity are indicated. Yellow is used to identify the HDACs activators, while light blue is used to identify the HDACs inhibitors. To see this illustration in color, the reader is referred to the web version of this article at www.lie bertpub.com/ars

11

physical changes, hypoxia, a typical feature of developing tumors that arise after the reduction of vascularization (33, 67), activates HDACs. Similarly, also the pH lowering, a common feature in tumors as consequence of an increased fermentative metabolism of glucose and accumulation of lactic acid in the space surrounding the tumor (20), induces HDACs activity. Likewise, several metabolites regulating HDACs activity are products of metabolic pathways that are deregulated in cancer. Tumors show increased levels of NADPH, an HDACs activator, in order to sustain several anabolic reactions; for example, nucleotides, glutathione, and, the most important, fatty acid synthesis (36, 79, 80). In addition, different intermediates of the mevalonate pathway (e.g., acetoacetyl-CoA, HMG-CoA), a metabolic pathway involved in cholesterol synthesis and protein prenylation and often highly activated in cancer (45), positively regulate HDACs. Equally, several intermediates of amino-acid degradation (acetyl-CoA, crotonyl-CoA, glutaryl-CoA, methylcrotonyl-CoA, methylmalonyl-CoA, and succinyl-CoA) that cancer cells use for fatty acid synthesis and other anabolic needs (15, 102, 135, 138, 174) activate HDACs. Interestingly, a correlation between HDACs activity and deregulated metabolism could be also observed for the negative regulators. Palmitolyl-CoA, for instance, being a substrate for the synthesis of sphingolipids, often increased in cancer, is generally rapidly used in tumors. Contrasting findings have been observed for sphingosine-1P. Nevertheless, this metabolic molecule has an important role in cancer proliferation and survival, and its balance is tightly regulated in tumors (123, 140). l-Carnitine, able to inhibit HDACs activity, is usually low in tumor cells, as they favor fatty acid synthesis over oxidation (2, 76). This cancer cell feature is also the reason, because usually in cancer there are low levels of bOHB, another HDACi. In fact, the formation of ketone bodies is usually repressed in cancer cells, as they favor lipid

12

synthesis over the ketogenic route (14, 69). Pyruvate and lactic acid, generated essentially through glycolysis, can also influence HDAC activity. However, while the inhibitory effect of pyruvate on HDACs is expected to be reduced in cancer, since it is rapidly transformed in lactic acid in glycolytic tumors (164), a dual role may be assigned to lactic acid. In fact, its ability to decrease extracellular pH on excretion, leading to a decrease of intracellular pH, may lead to a positive effect on HDAC activity. Conversely, it has been shown that at a high concentration, it may negatively affect HDACs activity. Such a dual role needs further investigation. The role of butyrate has been extensively discussed earlier; here, we only want to underline the protective effect of butyrate against colon cancer due to its ability to inhibit HDACs. Such an effect has been observed in vitro on colon cancer cells and in vivo through a high fiber diet (69). Altogether, these findings further suggest a connection between metabolism and HDACs, especially in cancer. Conclusion and Outlook

Here, we have discussed the cross-talk between classical HDACs and cell metabolism with particular emphasis regarding cancer cell metabolism. We have seen that metabolism affects HDACs activity through several intermediates of different metabolic pathways as well as through signaling pathways which are responsive to metabolic changes (Fig. 8). As a consequence, metabolism may control transcription and several other processes regulated by the HDACs activity on non-histone proteins. On the other hand, HDACs control cell

CHIARADONNA ET AL.

metabolism by deacetylating metabolic enzymes (Fig. 8). Such a relationship is clearly shown through the use of novel and more sophisticated technologies that are permitting the isolation of a high number of acetylated metabolic enzymes whose activity is changed on this post-translational modification. Interestingly, the acetylation state of these enzymes has been suggested to change depending on extracellular nutrient availability. In addition, recent reports, at least in rodent models, indicate that glucose homeostasis is actively regulated by HDACs as well as HDACs activity and expression is involved in metabolic changes, observed in different organisms such as rat and monkey, triggered by the diet (Fig. 8). Despite the mechanisms controlling the connection between acetylation/deacetylation and metabolism still remaining unclear, the identification of a physiological role of HDACs in controlling metabolism has further increased the interest for their role in pathological conditions, such as diabetes, obesity, and cancer, as they are considered possible therapeutic targets for these diseases. However, the use of pan-HDACs inhibitors in cell and mouse models as well as in phase I–III clinical trials for malignant diseases has clearly indicated two important points: First, the mechanism of action of HDACi is only partially related to altered gene expression, as all the inhibitors on average modulate 2% of total genes (56, 111), and they are more related to changes in nonhistone proteins via regulation at the post-translational modification level. Second, HDACi have varying antitumor activity, as their effects are influenced by the cell type, the dose, and the type of tumors. Both points may suggest that the

FIG. 8. Metabolism and HDACs: a novel cross-talk associated to physiological and pathological regulation of cell metabolism. As described extensively in the main text, several endogenous and exogenous stimuli regulate HDACs activity both positively and negatively ( + / - ). On the other hand, once activated, HDACs may regulate metabolism through their transcriptional regulatory function on histone proteins or through their deacetylase activity on nonhistone proteins. Both effects have been described in cell models and organismal models in either physiological or pathological conditions. Such an interconnection between HDACs and metabolism has been clearly shown by studying the in vivo effects of the HDACi that, among other effects, induce a large metabolic rearrangement. HDACi, histone deacetylase inhibitor.

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

different repertoire of non-histone protein targets in a tumor may be the reason for the variability of their efficacy. In this scenario, the identification of HDACs substrates will be increasingly interesting. Recent data have indicated metabolic enzymes as an important class of targets. In fact, it has been observed that virtually each enzyme of numerous metabolic pathways is modified by acetylation, as also shown by the large metabolic alterations observed on using HDACi treatment in cell models, indicating an important role of HDACs in their function (Fig. 8). Altogether, these findings have to be taken more into account while considering the revival of cancer metabolism as a key hallmark of tumorigenesis. In fact, the discovery that

13

different metabolic intermediates, highly produced by some of the pathways deregulated in cancer cells, such as fatty acid synthesis and pentose phosphate pathways, or chemical changes characterizing the tumors, such as hypoxia and low pH, may positively regulate HDAC activity (Fig. 9A) suggests a tight association between the two processes and probably nutrient availability. On the other hand, the ability of HDACs, as largely discussed throughout the text, to modulate the same metabolic pathways through their effect on oncosuppressors and oncogene activity implies the existence of a ‘‘positive feedback’’ for cancer in which metabolism and HDACs cross-talk to drive tumorigenesis. However, as anticipated throughout the review, several gaps in our

FIG. 9. HDACs and metabolism concur to the formation of a ‘‘positive feedback’’ for cancer. (A) The different findings collected in this review suggest the existence in cancer of a ‘‘feedback cycle’’ supporting tumorigenesis. In fact, highly produced metabolic intermediates (from fatty acid synthesis or Pentose Phosphate Pathway) and some chemical changes (hypoxia and low pH), both typical of the tumors, participate in HDAC activation. Then, HDACs, through their effect on oncogenes and oncosuppressors, lead to the stimulation of the same metabolic pathways, closing the circle. (B) In this scenario, in the future, several questions need to be addressed, and, in particular, it will be extremely interesting to discover and define the reciprocal influences along the tumorigenic process between these two cellular processes, especially on comparing the different metabolic cancer types in relation to the different HDACs classes. Major advances will sure serve as a warranty by the application of the metabolomics techniques. This will open new therapeutic applications not only in cancer but also in other metabolic-related diseases. To see this illustration in color, the reader is referred to the web version of this article at www.liebertpub.com/ars

14

knowledge regarding this ‘‘feedback’’ need to be filled. For instance, defining the relationship between the different metabolic profiles of the tumors, for example, glycolytic or glutaminolytic, and the specific class of HDACs active in the same type of the tumors will help delineate the reciprocal influences along the tumorigenic process (Fig. 9B). Another challenge for the future will be also to discover the pathological roles of individual HDACs in cancer metabolic alterations by using specific KO mice or cell lines that will also permit to identify HDAC isoform-specific targets (Fig. 9B). Similarly, it is currently almost unclear whether really cancer-specific metabolic intermediates may participate in HDAC activation in tumors. Altogether, these aims may find answers in the novel field of metabolomics, as the analysis of metabolites represents a sensitive measure of biological status in health or disease and may give fingerprints that are sometimes unique to every individual or every disease (Fig. 9B). In this regard, a systems approach is often useful to better understand the execution and control of complex biological functions (90). The first step of this approach generally requires a clear definition of the gene partners involved in specific modules (for instance, in our case, the regulation of a metabolic pathway by a given HDAC) (3). The availability of an interactome map of the human HDAC family (82) may enable a definition of the molecular network that underlies the function under investigation. The construction of a concept map following lines reported in the literature [e.g., Alzheimer Disease map (120)] brings a step further toward the construction of a dynamic molecular model, which generally involves a restricted number of molecular players—the ones involved in a more critical feature of the function under observation. Simulation analysis and experimental validation are finally able to shed light on the design principles that are crucial for the establishment of the physiological behavior. The replication of this procedure to several aspects of HDAC regulation of cancer cell metabolism may help identify more effective (possible combinatorial) drug targets. Acknowledgments

The authors apologize to the many authors whose work they were not able to cite given space constraints. Research in Chiaradonna laboratory has been supported by grants to F.C. from the Italian Government (FAR). L.A. and F.C. have been partially supported by SysBioNet, an MIUR grant for the Italian Roadmap of ESFRI Infrastructures. R.P. and G.V. have been supported by fellowships of SysBioNet. C.C. has been supported by a grant from MIUR (FIRB, 2008). The authors wish to thank Neil Campbell for English editing.

CHIARADONNA ET AL.

4. 5.

6.

7.

8.

9.

10.

11.

12. 13. 14.

15.

References

1. Aagaard-Tillery KM, Grove K, Bishop J, Ke X, Fu Q, McKnight R, and Lane RH. Developmental origins of disease and determinants of chromatin structure: maternal diet modifies the primate fetal epigenome. J Mol Endocrinol 41: 91–102, 2008. 2. Abramson HN. The lipogenesis pathway as a cancer target. J Med Chem 54: 5615–5638, 2011. 3. Alberghina L, Gaglio D, Gelfi C, Moresco RM, Mauri G, Bertolazzi P, Messa C, Gilardi MC, Chiaradonna F, and Vanoni M. Cancer cell growth and survival as a systemlevel property sustained by enhanced glycolysis and mi-

16.

17.

18.

tochondrial metabolic remodeling. Front Physiol 3: 362, 2012. Alirol E and Martinou JC. Mitochondria and cancer: is there a morphological connection? Oncogene 25: 4706– 4716, 2006. Amoedo ND, Rodrigues MF, Pezzuto P, Galina A, da Costa RM, de Almeida FC, El-Bacha T, and Rumjanek FD. Energy metabolism in H460 lung cancer cells: effects of histone deacetylase inhibitors. PLoS One 6: e22264, 2011. Austin S and St.-Pierre J. PGC1alpha and mitochondrial metabolism—emerging concepts and relevance in ageing and neurodegenerative disorders. J Cell Sci 125: 4963– 4971, 2012. Backs J, Worst BC, Lehmann LH, Patrick DM, Jebessa Z, Kreusser MM, Sun Q, Chen L, Heft C, Katus HA, and Olson EN. Selective repression of MEF2 activity by PKAdependent proteolysis of HDAC4. J Cell Biol 195: 403– 415, 2011. Beloueche-Babari M, Arunan V, Troy H, te Poele RH, te Fong AC, Jackson LE, Payne GS, Griffiths JR, Judson IR, Workman P, Leach MO, and Chung YL. Histone deacetylase inhibition increases levels of choline kinase alpha and phosphocholine facilitating noninvasive imaging in human cancers. Cancer Res 72: 990–1000, 2012. Bensaad K, Tsuruta A, Selak MA, Vidal MN, Nakano K, Bartrons R, Gottlieb E, and Vousden KH. TIGAR, a p53inducible regulator of glycolysis and apoptosis. Cell 126: 107–120, 2006. Blattler SM, Verdeguer F, Liesa M, Cunningham JT, Vogel RO, Chim H, Liu H, Romanino K, Shirihai OS, Vazquez F, Ruegg MA, Shi Y, and Puigserver P. Defective mitochondrial morphology and bioenergetic function in mice lacking the transcription factor Yin Yang 1 in skeletal muscle. Mol Cell Biol 32: 3333–3346, 2012. Bolden JE, Shi W, Jankowski K, Kan CY, Cluse L, Martin BP, MacKenzie KL, Smyth GK, and Johnstone RW. HDAC inhibitors induce tumor-cell-selective pro-apoptotic transcriptional responses. Cell Death Dis 4: e519, 2013. Boxer LM and Dang CV. Translocations involving c-myc and c-myc function. Oncogene 20: 5595–5610, 2001. Brahimi-Horn MC, Chiche J, and Pouyssegur J. Hypoxia signalling controls metabolic demand. Curr Opin Cell Biol 19: 223–229, 2007. Buzzai M, Bauer DE, Jones RG, Deberardinis RJ, Hatzivassiliou G, Elstrom RL, and Thompson CB. The glucose dependence of Akt-transformed cells can be reversed by pharmacologic activation of fatty acid beta-oxidation. Oncogene 24: 4165–4173, 2005. Cavuoto P and Fenech MF. A review of methionine dependency and the role of methionine restriction in cancer growth control and life-span extension. Cancer Treat Rev 38: 726–736, 2012. Chang CW, Lee L, Yu D, Dao K, Bossuyt J, and Bers DM. Acute beta-adrenergic activation triggers nuclear import of histone deacetylase 5 and delays G(q)-induced transcriptional activation. J Biol Chem 288: 192–204, 2013. Chang S, McKinsey TA, Zhang CL, Richardson JA, Hill JA, and Olson EN. Histone deacetylases 5 and 9 govern responsiveness of the heart to a subset of stress signals and play redundant roles in heart development. Mol Cell Biol 24: 8467–8476, 2004. Cheng J, Wang D, Wang Z, and Yeh ET. SENP1 enhances androgen receptor-dependent transcription through

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

19.

20.

21. 22.

23. 24. 25. 26.

27.

28. 29.

30. 31.

32.

33. 34.

35.

desumoylation of histone deacetylase 1. Mol Cell Biol 24: 6021–6028, 2004. Chiaradonna F, Gaglio D, Vanoni M, and Alberghina L. Expression of transforming K-Ras oncogene affects mitochondrial function and morphology in mouse fibroblasts. Biochim Biophys Acta 1757: 1338–1356, 2006. Chiaradonna F, Moresco RM, Airoldi C, Gaglio D, Palorini R, Nicotra F, Messa C, and Alberghina L. From cancer metabolism to new biomarkers and drug targets. Biotechnol Adv 30: 30–51, 2012. Chou CW, Wu MS, Huang WC, and Chen CC. HDAC inhibition decreases the expression of EGFR in colorectal cancer cells. PLoS One 6: e18087, 2011. Choudhary C, Kumar C, Gnad F, Nielsen ML, Rehman M, Walther TC, Olsen JV, and Mann M. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science 325: 834–840, 2009. Cohen I, Poreba E, Kamieniarz K, and Schneider R. Histone modifiers in cancer: friends or foes? Genes Cancer 2: 631–647, 2011. Conte M and Altucci L. Molecular pathways: the complexity of the epigenome in cancer and recent clinical advances. Clin Cancer Res 18: 5526–5534, 2011. Cousens LS, Gallwitz D, and Alberts BM. Different accessibilities in chromatin to histone acetylase. J Biol Chem 254: 1716–1723, 1979. Cunningham JT, Rodgers JT, Arlow DH, Vazquez F, Mootha VK, and Puigserver P. mTOR controls mitochondrial oxidative function through a YY1-PGC-1alpha transcriptional complex. Nature 450: 736–740, 2007. Czubryt MP, McAnally J, Fishman GI, and Olson EN. Regulation of peroxisome proliferator-activated receptor gamma coactivator 1 alpha (PGC-1 alpha) and mitochondrial function by MEF2 and HDAC5. Proc Natl Acad Sci U S A 100: 1711–1716, 2003. Dang CV, Le A, and Gao P. MYC-induced cancer cell energy metabolism and therapeutic opportunities. Clin Cancer Res 15: 6479–6483, 2009. DeAngelis RA, Markiewski MM, Taub R, and Lambris JD. A high-fat diet impairs liver regeneration in C57BL/6 mice through overexpression of the NF-kappaB inhibitor, IkappaBalpha. Hepatology 42: 1148–1157, 2005. DeBerardinis RJ and Cheng T. Q’s next: the diverse functions of glutamine in metabolism, cell biology and cancer. Oncogene 29: 313–324, 2010. DeBerardinis RJ, Lum JJ, Hatzivassiliou G, and Thompson CB. The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab 7: 11–20, 2008. Demaria M, Giorgi C, Lebiedzinska M, Esposito G, D’Angeli L, Bartoli A, Gough DJ, Turkson J, Levy DE, Watson CJ, Wieckowski MR, Provero P, Pinton P, and Poli V. A STAT3-mediated metabolic switch is involved in tumour transformation and STAT3 addiction. Aging (Albany NY) 2: 823–842, 2010. Dewhirst MW, Cao Y, and Moeller B. Cycling hypoxia and free radicals regulate angiogenesis and radiotherapy response. Nat Rev Cancer 8: 425–437, 2008. Donohoe DR, Collins LB, Wali A, Bigler R, Sun W, and Bultman SJ. The Warburg effect dictates the mechanism of butyrate-mediated histone acetylation and cell proliferation. Mol Cell 48: 612–626, 2012. Dowling DP, Gattis SG, Fierke CA, and Christianson DW. Structures of metal-substituted human histone deacetylase

36.

37.

38.

39.

40.

41.

42.

43.

44.

45.

46.

47.

48.

15

8 provide mechanistic inferences on biological function. Biochemistry 49: 5048–5056, 2010. Duvel K, Yecies JL, Menon S, Raman P, Lipovsky AI, Souza AL, Triantafellow E, Ma Q, Gorski R, Cleaver S, Vander Heiden MG, MacKeigan JP, Finan PM, Clish CB, Murphy LO, and Manning BD. Activation of a metabolic gene regulatory network downstream of mTOR complex 1. Mol Cell 39: 171–183, 2010. Elstrom RL, Bauer DE, Buzzai M, Karnauskas R, Harris MH, Plas DR, Zhuang H, Cinalli RM, Alavi A, Rudin CM, and Thompson CB. Akt stimulates aerobic glycolysis in cancer cells. Cancer Res 64: 3892–3899, 2004. Faiola F, Liu X, Lo S, Pan S, Zhang K, Lymar E, Farina A, and Martinez E. Dual regulation of c-Myc by p300 via acetylation-dependent control of Myc protein turnover and coactivation of Myc-induced transcription. Mol Cell Biol 25: 10220–10234, 2005. Fang X, Yu SX, Lu Y, Bast RC, Jr., Woodgett JR, and Mills GB. Phosphorylation and inactivation of glycogen synthase kinase 3 by protein kinase A. Proc Natl Acad Sci U S A 97: 11960–11965, 2000. Faubert B, Boily G, Izreig S, Griss T, Samborska B, Dong Z, Dupuy F, Chambers C, Fuerth BJ, Viollet B, Mamer OA, Avizonis D, DeBerardinis RJ, Siegel PM, and Jones RG. AMPK is a negative regulator of the Warburg effect and suppresses tumor growth in vivo. Cell Metab 17: 113– 124, 2013. Ferber EC, Peck B, Delpuech O, Bell GP, East P, and Schulze A. FOXO3a regulates reactive oxygen metabolism by inhibiting mitochondrial gene expression. Cell Death Differ 19: 968–979, 2011. Fischer DD, Cai R, Bhatia U, Asselbergs FA, Song C, Terry R, Trogani N, Widmer R, Atadja P, and Cohen D. Isolation and characterization of a novel class II histone deacetylase, HDAC10. J Biol Chem 277: 6656–6666, 2002. Fischle W, Dequiedt F, Fillion M, Hendzel MJ, Voelter W, and Verdin E. Human HDAC7 histone deacetylase activity is associated with HDAC3 in vivo. J Biol Chem 276: 35826–35835, 2001. Frank SR, Parisi T, Taubert S, Fernandez P, Fuchs M, Chan HM, Livingston DM, and Amati B. MYC recruits the TIP60 histone acetyltransferase complex to chromatin. EMBO Rep 4: 575–580, 2003. Freed-Pastor WA, Mizuno H, Zhao X, Langerod A, Moon SH, Rodriguez-Barrueco R, Barsotti A, Chicas A, Li W, Polotskaia A, Bissell MJ, Osborne TF, Tian B, Lowe SW, Silva JM, Borresen-Dale AL, Levine AJ, Bargonetti J, and Prives C. Mutant p53 disrupts mammary tissue architecture via the mevalonate pathway. Cell 148: 244–258, 2012. Galmozzi A, Mitro N, Ferrari A, Gers E, Gilardi F, Godio C, Cermenati G, Gualerzi A, Donetti E, Rotili D, Valente S, Guerrini U, Caruso D, Mai A, Saez E, De Fabiani E, and Crestani M. Inhibition of class I histone deacetylases unveils a mitochondrial signature and enhances oxidative metabolism in skeletal muscle and adipose tissue. Diabetes 62: 732–742, 2012. Gan YH and Zhang S. PTEN/AKT pathway involved in histone deacetylases inhibitor induced cell growth inhibition and apoptosis of oral squamous cell carcinoma cells. Oral Oncol 45: e150–e154, 2009. Gao C, Dimitrov T, Yong KJ, Tatetsu H, Jeong HW, Luo HR, Bradner JE, Tenen DG, and Chai L. Targeting transcription factor SALL4 in acute myeloid leukemia by

16

49.

50.

51.

52. 53. 54.

55.

56.

57.

58. 59.

60. 61.

62. 63.

64.

65.

CHIARADONNA ET AL.

interrupting its interaction with an epigenetic complex. Blood 121: 1413–1421, 2013. Gao L, Cueto MA, Asselbergs F, and Atadja P. Cloning and functional characterization of HDAC11, a novel member of the human histone deacetylase family. J Biol Chem 277: 25748–25755, 2002. Gelmetti V, Zhang J, Fanelli M, Minucci S, Pelicci PG, and Lazar MA. Aberrant recruitment of the nuclear receptor corepressor-histone deacetylase complex by the acute myeloid leukemia fusion partner ETO. Mol Cell Biol 18: 7185–7191, 1998. Geng H, Harvey CT, Pittsenbarger J, Liu Q, Beer TM, Xue C, and Qian DZ. HDAC4 protein regulates HIF1alpha protein lysine acetylation and cancer cell response to hypoxia. J Biol Chem 286: 38095–38102, 2011. Georgescu MM. PTEN tumor suppressor network in PI3KAkt pathway control. Genes Cancer 1: 1170–1177, 2010. Glozak MA and Seto E. Histone deacetylases and cancer. Oncogene 26: 5420–5432, 2007. Gordon S, Akopyan G, Garban H, and Bonavida B. Transcription factor YY1: structure, function, and therapeutic implications in cancer biology. Oncogene 25: 1125–1142, 2006. Gough DJ, Corlett A, Schlessinger K, Wegrzyn J, Larner AC, and Levy DE. Mitochondrial STAT3 supports Rasdependent oncogenic transformation. Science 324: 1713– 1716, 2009. Gray SG, Qian CN, Furge K, Guo X, and Teh BT. Microarray profiling of the effects of histone deacetylase inhibitors on gene expression in cancer cell lines. Int J Oncol 24: 773–795, 2004. Gregoretti IV, Lee YM, and Goodson HV. Molecular evolution of the histone deacetylase family: functional implications of phylogenetic analysis. J Mol Biol 338: 17– 31, 2004. Gross DN, van den Heuvel AP, and Birnbaum MJ. The role of FoxO in the regulation of metabolism. Oncogene 27: 2320–2336, 2008. Grozinger CM and Schreiber SL. Regulation of histone deacetylase 4 and 5 and transcriptional activity by 14-3-3dependent cellular localization. Proc Natl Acad Sci U S A 97: 7835–7840, 2000. Gu W and Roeder RG. Activation of p53 sequencespecific DNA binding by acetylation of the p53 C-terminal domain. Cell 90: 595–606, 1997. Gupta M, Han JJ, Stenson M, Wellik L, and Witzig TE. Regulation of STAT3 by histone deacetylase-3 in diffuse large B-cell lymphoma: implications for therapy. Leukemia 26: 1356–1364, 2012. Haberland M, Mokalled MH, Montgomery RL, and Olson EN. Epigenetic control of skull morphogenesis by histone deacetylase 8. Genes Dev 23: 1625–1630, 2009. Haberland M, Montgomery RL, and Olson EN. The many roles of histone deacetylases in development and physiology: implications for disease and therapy. Nat Rev Genet 10: 32–42, 2009. Hait NC, Allegood J, Maceyka M, Strub GM, Harikumar KB, Singh SK, Luo C, Marmorstein R, Kordula T, Milstien S, and Spiegel S. Regulation of histone acetylation in the nucleus by sphingosine-1-phosphate. Science 325: 1254–1257, 2009. Hardie DG and Alessi DR. LKB1 and AMPK and the cancer-metabolism link—ten years after. BMC Biol 11: 36, 2013.

66. Harms KL and Chen X. Histone deacetylase 2 modulates p53 transcriptional activities through regulation of p53DNA binding activity. Cancer Res 67: 3145–3152, 2007. 67. Harris AL. Hypoxia—a key regulatory factor in tumour growth. Nat Rev Cancer 2: 38–47, 2002. 68. Harrison BC, Huynh K, Lundgaard GL, Helmke SM, Perryman MB, and McKinsey TA. Protein kinase Crelated kinase targets nuclear localization signals in a subset of class IIa histone deacetylases. FEBS Lett 584: 1103–1110, 2010. 69. Hassig CA, Tong JK, and Schreiber SL. Fiber-derived butyrate and the prevention of colon cancer. Chem Biol 4: 783–789, 1997. 70. Hebert AS, Dittenhafer-Reed KE, Yu W, Bailey DJ, Selen ES, Boersma MD, Carson JJ, Tonelli M, Balloon AJ, Higbee AJ, Westphall MS, Pagliarini DJ, Prolla TA, Assadi-Porter F, Roy S, Denu JM, and Coon JJ. Calorie restriction and SIRT3 trigger global reprogramming of the mitochondrial protein acetylome. Mol Cell 49: 186–199, 2013. 71. Heruth DP, Zirnstein GW, Bradley JF, and Rothberg PG. Sodium butyrate causes an increase in the block to transcriptional elongation in the c-myc gene in SW837 rectal carcinoma cells. J Biol Chem 268: 20466–20472, 1993. 72. Hervouet E, Demont J, Pecina P, Vojtiskova A, Houstek J, Simonnet H, and Godinot C. A new role for the von Hippel-Lindau tumor suppressor protein: stimulation of mitochondrial oxidative phosphorylation complex biogenesis. Carcinogenesis 26: 531–539, 2005. 73. Huang H, Liu N, Guo H, Liao S, Li X, Yang C, Liu S, Song W, Liu C, Guan L, Li B, Xu L, Zhang C, Wang X, Dou QP, and Liu J. L-carnitine is an endogenous HDAC inhibitor selectively inhibiting cancer cell growth in vivo and in vitro. PLoS One 7: e49062, 2012. 74. Ismail HM, Myronova O, Tsuchiya Y, Niewiarowski A, Tsaneva I, and Gout I. Identification of the general transcription factor Yin Yang 1 as a novel and specific binding partner for S6 kinase 2. Cell Signal 25: 1054–1063, 2013. 75. Ito A, Kawaguchi Y, Lai CH, Kovacs JJ, Higashimoto Y, Appella E, and Yao TP. MDM2-HDAC1-mediated deacetylation of p53 is required for its degradation. EMBO J 21: 6236–6245, 2002. 76. Jain M, Nilsson R, Sharma S, Madhusudhan N, Kitami T, Souza AL, Kafri R, Kirschner MW, Clish CB, and Mootha VK. Metabolite profiling identifies a key role for glycine in rapid cancer cell proliferation. Science 336: 1040–1044, 2012. 77. Jaitovich A and Bertorello AM. Intracellular sodium sensing: SIK1 network, hormone action and high blood pressure. Biochim Biophys Acta 1802: 1140–1149, 2010. 78. Jensen J, Rustad PI, Kolnes AJ, and Lai YC. The role of skeletal muscle glycogen breakdown for regulation of insulin sensitivity by exercise. Front Physiol 2: 112, 2011. 79. Jiang P, Du W, Mancuso A, Wellen KE, and Yang X. Reciprocal regulation of p53 and malic enzymes modulates metabolism and senescence. Nature 493: 689–693, 2013. 80. Jiang P, Du W, Wang X, Mancuso A, Gao X, Wu M, and Yang X. p53 regulates biosynthesis through direct inactivation of glucose-6-phosphate dehydrogenase. Nat Cell Biol 13: 310–316, 2011. 81. Joensuu T and Mester J. Inhibition of cell cycle progression by sodium butyrate in normal rat kidney fibroblasts is altered by expression of the adenovirus 5 early 1A gene. Biosci Rep 14: 291–300, 1994.

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

82. Joshi P, Greco TM, Guise AJ, Luo Y, Yu F, Nesvizhskii AI, and Cristea IM. The functional interactome landscape of the human histone deacetylase family. Mol Syst Biol 9: 672, 2013. 83. Juan LJ, Shia WJ, Chen MH, Yang WM, Seto E, Lin YS, and Wu CW. Histone deacetylases specifically downregulate p53-dependent gene activation. J Biol Chem 275: 20436–20443, 2000. 84. Kaelin WG, Jr., and McKnight SL. Influence of metabolism on epigenetics and disease. Cell 153: 56–69, 2013. 85. Kato H, Tamamizu-Kato S, and Shibasaki F. Histone deacetylase 7 associates with hypoxia-inducible factor 1alpha and increases transcriptional activity. J Biol Chem 279: 41966–41974, 2004. 86. Kelly RD and Cowley SM. The physiological roles of histone deacetylase (HDAC) 1 and 2: complex co-stars with multiple leading parts. Biochem Soc Trans 41: 741– 749, 2013. 87. Kim MS, Kwon HJ, Lee YM, Baek JH, Jang JE, Lee SW, Moon EJ, Kim HS, Lee SK, Chung HY, Kim CW, and Kim KW. Histone deacetylases induce angiogenesis by negative regulation of tumor suppressor genes. Nat Med 7: 437–443, 2001. 88. Kim SC, Sprung R, Chen Y, Xu Y, Ball H, Pei J, Cheng T, Kho Y, Xiao H, Xiao L, Grishin NV, White M, Yang XJ, and Zhao Y. Substrate and functional diversity of lysine acetylation revealed by a proteomics survey. Mol Cell 23: 607–618, 2006. 89. Kirsh O, Seeler JS, Pichler A, Gast A, Muller S, Miska E, Mathieu M, Harel-Bellan A, Kouzarides T, Melchior F, and Dejean A. The SUMO E3 ligase RanBP2 promotes modification of the HDAC4 deacetylase. EMBO J 21: 2682–2691, 2002. 90. Kitano H. Systems biology: a brief overview. Science 295: 1662–1664, 2002. 91. Knutson SK, Chyla BJ, Amann JM, Bhaskara S, Huppert SS, and Hiebert SW. Liver-specific deletion of histone deacetylase 3 disrupts metabolic transcriptional networks. EMBO J 27: 1017–1028, 2008. 92. Kong X, Lin Z, Liang D, Fath D, Sang N, and Caro J. Histone deacetylase inhibitors induce VHL and ubiquitinindependent proteasomal degradation of hypoxia-inducible factor 1alpha. Mol Cell Biol 26: 2019–2028, 2006. 93. Kops GJ, Dansen TB, Polderman PE, Saarloos I, Wirtz KW, Coffer PJ, Huang TT, Bos JL, Medema RH, and Burgering BM. Forkhead transcription factor FOXO3a protects quiescent cells from oxidative stress. Nature 419: 316–321, 2002. 94. Kou XX, Hao T, Meng Z, Zhou YH, and Gan YH. Acetylated Sp1 inhibits PTEN expression through binding to PTEN core promoter and recruitment of HDAC1 and promotes cancer cell migration and invasion. Carcinogenesis 34: 58–67, 2013. 95. Kovacs JJ, Murphy PJ, Gaillard S, Zhao X, Wu JT, Nicchitta CV, Yoshida M, Toft DO, Pratt WB, and Yao TP. HDAC6 regulates Hsp90 acetylation and chaperonedependent activation of glucocorticoid receptor. Mol Cell 18: 601–607, 2005. 96. Kretzner L, Scuto A, Dino PM, Kowolik CM, Wu J, Ventura P, Jove R, Forman SJ, Yen Y, and Kirschbaum MH. Combining histone deacetylase inhibitor vorinostat with aurora kinase inhibitors enhances lymphoma cell killing with repression of c-Myc, hTERT, and microRNA levels. Cancer Res 71: 3912–3920, 2011.

17

97. Kuhajda FP. Fatty-acid synthase and human cancer: new perspectives on its role in tumor biology. Nutrition 16: 202–208, 2000. 98. Kurland JF and Tansey WP. Myc-mediated transcriptional repression by recruitment of histone deacetylase. Cancer Res 68: 3624–3629, 2008. 99. Lagger G, O’Carroll D, Rembold M, Khier H, Tischler J, Weitzer G, Schuettengruber B, Hauser C, Brunmeir R, Jenuwein T, and Seiser C. Essential function of histone deacetylase 1 in proliferation control and CDK inhibitor repression. EMBO J 21: 2672–2681, 2002. 100. Laplante M and Sabatini DM. Regulation of mTORC1 and its impact on gene expression at a glance. J Cell Sci 126: 1713–1719, 2013. 101. Latham T, Mackay L, Sproul D, Karim M, Culley J, Harrison DJ, Hayward L, Langridge-Smith P, Gilbert N, and Ramsahoye BH. Lactate, a product of glycolytic metabolism, inhibits histone deacetylase activity and promotes changes in gene expression. Nucleic Acids Res 40: 4794–4803, 2012. 102. Lazo PA. Amino acids and glucose utilization by different metabolic pathways in ascites-tumour cells. Eur J Biochem 117: 19–25, 1981. 103. Lee H, Rezai-Zadeh N, and Seto E. Negative regulation of histone deacetylase 8 activity by cyclic AMP-dependent protein kinase A. Mol Cell Biol 24: 765–773, 2004. 104. Li DQ, Pakala SB, Nair SS, Eswaran J, and Kumar R. Metastasis-associated protein 1/nucleosome remodeling and histone deacetylase complex in cancer. Cancer Res 72: 387–394, 2012. 105. Liang J and Mills GB. AMPK: a contextual oncogene or tumor suppressor? Cancer Res 73: 2929–2935, 2013. 106. Lin R, Zhou X, Huang W, Zhao D, Lv L, Xiong Y, Guan KL, and Lei QY. Acetylation control of cancer cell metabolism. Curr Pharm Des 13: 1–7, 2013. 107. Lin YY, Kiihl S, Suhail Y, Liu SY, Chou YH, Kuang Z, Lu JY, Khor CN, Lin CL, Bader JS, Irizarry R, and Boeke JD. Functional dissection of lysine deacetylases reveals that HDAC1 and p300 regulate AMPK. Nature 482: 251– 255, 2012. 108. Luo J, Su F, Chen D, Shiloh A, and Gu W. Deacetylation of p53 modulates its effect on cell growth and apoptosis. Nature 408: 377–381, 2000. 109. Lv L, Li D, Zhao D, Lin R, Chu Y, Zhang H, Zha Z, Liu Y, Li Z, Xu Y, Wang G, Huang Y, Xiong Y, Guan KL, and Lei QY. Acetylation targets the M2 isoform of pyruvate kinase for degradation through chaperone-mediated autophagy and promotes tumor growth. Mol Cell 42: 719–730, 2011. 110. Marks PA. The clinical development of histone deacetylase inhibitors as targeted anticancer drugs. Expert Opin Investig Drugs 19: 1049–1066, 2010. 111. Marks PA, Richon VM, Miller T, and Kelly WK. Histone deacetylase inhibitors. Adv Cancer Res 91: 137–168, 2004. 112. Marshall GM, Gherardi S, Xu N, Neiron Z, Trahair T, Scarlett CJ, Chang DK, Liu PY, Jankowski K, Iraci N, Haber M, Norris MD, Keating J, Sekyere E, Jonquieres G, Stossi F, Katzenellenbogen BS, Biankin AV, Perini G, and Liu T. Transcriptional upregulation of histone deacetylase 2 promotes Myc-induced oncogenic effects. Oncogene 29: 5957–5968, 2010. 113. Matoba S, Kang JG, Patino WD, Wragg A, Boehm M, Gavrilova O, Hurley PJ, Bunz F, and Hwang PM. p53 regulates mitochondrial respiration. Science 312: 1650– 1653, 2006.

18

114. McBrian MA, Behbahan IS, Ferrari R, Su T, Huang TW, Li K, Hong CS, Christofk HR, Vogelauer M, Seligson DB, and Kurdistani SK. Histone acetylation regulates intracellular pH. Mol Cell 49: 310–321, 2013. 115. McGee SL, van Denderen BJ, Howlett KF, Mollica J, Schertzer JD, Kemp BE, and Hargreaves M. AMP-activated protein kinase regulates GLUT4 transcription by phosphorylating histone deacetylase 5. Diabetes 57: 860–867, 2008. 116. Michalopoulos GK. Liver regeneration after partial hepatectomy: critical analysis of mechanistic dilemmas. Am J Pathol 176: 2–13, 2010. 117. Mihaylova MM and Shaw RJ. Metabolic reprogramming by class I and II histone deacetylases. Trends Endocrinol Metab 24: 48–57, 2013. 118. Mihaylova MM, Vasquez DS, Ravnskjaer K, Denechaud PD, Yu RT, Alvarez JG, Downes M, Evans RM, Montminy M, and Shaw RJ. Class IIa histone deacetylases are hormone-activated regulators of FOXO and mammalian glucose homeostasis. Cell 145: 607–621, 2011. 119. Minucci S and Pelicci PG. Histone deacetylase inhibitors and the promise of epigenetic (and more) treatments for cancer. Nat Rev Cancer 6: 38–51, 2006. 120. Mizuno S, Iijima R, Ogishima S, Kikuchi M, Matsuoka Y, Ghosh S, Miyamoto T, Miyashita A, Kuwano R, and Tanaka H. AlzPathway: a comprehensive map of signaling pathways of Alzheimer’s disease. BMC Syst Biol 6: 52, 2012. 121. Montgomery RL, Davis CA, Potthoff MJ, Haberland M, Fielitz J, Qi X, Hill JA, Richardson JA, and Olson EN. Histone deacetylases 1 and 2 redundantly regulate cardiac morphogenesis, growth, and contractility. Genes Dev 21: 1790–1802, 2007. 122. Montgomery RL, Potthoff MJ, Haberland M, Qi X, Matsuzaki S, Humphries KM, Richardson JA, Bassel-Duby R, and Olson EN. Maintenance of cardiac energy metabolism by histone deacetylase 3 in mice. J Clin Invest 118: 3588– 3597, 2008. 123. Morad SA and Cabot MC. Ceramide-orchestrated signalling in cancer cells. Nat Rev Cancer 13: 51–65, 2013. 124. Mosley AL and Ozcan S. The pancreatic duodenal homeobox-1 protein (Pdx-1) interacts with histone deacetylases Hdac-1 and Hdac-2 on low levels of glucose. J Biol Chem 279: 54241–54247, 2004. 125. Nishino TG, Miyazaki M, Hoshino H, Miwa Y, Horinouchi S, and Yoshida M. 14-3-3 regulates the nuclear import of class IIa histone deacetylases. Biochem Biophys Res Commun 377: 852–856, 2008. 126. Noro R, Gemma A, Miyanaga A, Kosaihira S, Minegishi Y, Nara M, Kokubo Y, Seike M, Kataoka K, Matsuda K, Okano T, Yoshimura A, and Kudoh S. PTEN inactivation in lung cancer cells and the effect of its recovery on treatment with epidermal growth factor receptor tyrosine kinase inhibitors. Int J Oncol 31: 1157–1163, 2007. 127. Ojuka EO, Goyaram V, and Smith JA. The role of CaMKII in regulating GLUT4 expression in skeletal muscle. Am J Physiol Endocrinol Metab 303: E322–E331, 2012. 128. Okumura K, Mendoza M, Bachoo RM, DePinho RA, Cavenee WK, and Furnari FB. PCAF modulates PTEN activity. J Biol Chem 281: 26562–26568, 2006. 129. Osoata GO, Yamamura S, Ito M, Vuppusetty C, Adcock IM, Barnes PJ, and Ito K. Nitration of distinct tyrosine residues causes inactivation of histone deacetylase 2. Biochem Biophys Res Commun 384: 366–371, 2009. 130. Ouyang H, Ali YO, Ravichandran M, Dong A, Qiu W, MacKenzie F, Dhe-Paganon S, Arrowsmith CH, and Zhai

CHIARADONNA ET AL.

131.

132.

133.

134. 135.

136.

137. 138. 139.

140.

141. 142.

143.

144. 145. 146.

RG. Protein aggregates are recruited to aggresome by histone deacetylase 6 via unanchored ubiquitin C termini. J Biol Chem 287: 2317–2327, 2011. Paik JH, Kollipara R, Chu G, Ji H, Xiao Y, Ding Z, Miao L, Tothova Z, Horner JW, Carrasco DR, Jiang S, Gilliland DG, Chin L, Wong WH, Castrillon DH, and DePinho RA. FoxOs are lineage-restricted redundant tumor suppressors and regulate endothelial cell homeostasis. Cell 128: 309– 323, 2007. Pan L, Lu J, Wang X, Han L, Zhang Y, Han S, and Huang B. Histone deacetylase inhibitor trichostatin a potentiates doxorubicin-induced apoptosis by up-regulating PTEN expression. Cancer 109: 1676–1688, 2007. Pasqualucci L, Dominguez-Sola D, Chiarenza A, Fabbri G, Grunn A, Trifonov V, Kasper LH, Lerach S, Tang H, Ma J, Rossi D, Chadburn A, Murty VV, Mullighan CG, Gaidano G, Rabadan R, Brindle PK, and Dalla-Favera R. Inactivating mutations of acetyltransferase genes in B-cell lymphoma. Nature 471: 189–195, 2011. Perez-Perri JI, Acevedo JM, and Wappner P. Epigenetics: new questions on the response to hypoxia. Int J Mol Sci 12: 4705–4721, 2011. Perroud B, Lee J, Valkova N, Dhirapong A, Lin PY, Fiehn O, Kultz D, and Weiss RH. Pathway analysis of kidney cancer using proteomics and metabolic profiling. Mol Cancer 5: 64, 2006. Petrie K, Guidez F, Howell L, Healy L, Waxman S, Greaves M, and Zelent A. The histone deacetylase 9 gene encodes multiple protein isoforms. J Biol Chem 278: 16059–16072, 2003. Pham TX and Lee J. Dietary regulation of histone acetylases and deacetylases for the prevention of metabolic diseases. Nutrients 4: 1868–1886, 2012. Platten M, Wick W, and Van den Eynde BJ. Tryptophan catabolism in cancer: beyond IDO and tryptophan depletion. Cancer Res 72: 5435–5440, 2012. Pluemsampant S, Safronova OS, Nakahama K, and Morita I. Protein kinase CK2 is a key activator of histone deacetylase in hypoxia-associated tumors. Int J Cancer 122: 333–341, 2008. Ponnusamy S, Meyers-Needham M, Senkal CE, Saddoughi SA, Sentelle D, Selvam SP, Salas A, and Ogretmen B. Sphingolipids and cancer: ceramide and sphingosine-1phosphate in the regulation of cell death and drug resistance. Future Oncol 6: 1603–1624, 2012. Prins ML. Cerebral metabolic adaptation and ketone metabolism after brain injury. J Cereb Blood Flow Metab 28: 1–16, 2008. Pugacheva EN, Jablonski SA, Hartman TR, Henske EP, and Golemis EA. HEF1-dependent Aurora A activation induces disassembly of the primary cilium. Cell 129: 1351–1363, 2007. Rodgers JT, Lerin C, Haas W, Gygi SP, Spiegelman BM, and Puigserver P. Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature 434: 113–118, 2005. Ropero S and Esteller M. The role of histone deacetylases (HDACs) in human cancer. Mol Oncol 1: 19–25, 2007. Saha RN and Pahan K. HATs and HDACs in neurodegeneration: a tale of disconcerted acetylation homeostasis. Cell Death Differ 13: 539–550, 2006. Sahlin K, Harris RC, Nylind B, and Hultman E. Lactate content and pH in muscle obtained after dynamic exercise. Pflugers Arch 367: 143–149, 1976.

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

147. Schulz TJ, Thierbach R, Voigt A, Drewes G, Mietzner B, Steinberg P, Pfeiffer AF, and Ristow M. Induction of oxidative metabolism by mitochondrial frataxin inhibits cancer growth: Otto Warburg revisited. J Biol Chem 281: 977–981, 2006. 148. Schwer B and Verdin E. Conserved metabolic regulatory functions of sirtuins. Cell Metab 7: 104–112, 2008. 149. Scott FL, Fuchs GJ, Boyd SE, Denault JB, Hawkins CJ, Dequiedt F, and Salvesen GS. Caspase-8 cleaves histone deacetylase 7 and abolishes its transcription repressor function. J Biol Chem 283: 19499–19510, 2008. 150. Segre CV and Chiocca S. Regulating the regulators: the post-translational code of class I HDAC1 and HDAC2. J Biomed Biotechnol 2011: 690848, 2011. 151. Semenza GL. Targeting HIF-1 for cancer therapy. Nat Rev Cancer 3: 721–732, 2003. 152. Semenza GL. Defining the role of hypoxia-inducible factor 1 in cancer biology and therapeutics. Oncogene 29: 625–634, 2010. 153. Sha Y and Marshall HE. S-nitrosylation in the regulation of gene transcription. Biochim Biophys Acta 1820: 701– 711, 2012. 154. Shimazu T, Hirschey MD, Newman J, He W, Shirakawa K, Le Moan N, Grueter CA, Lim H, Saunders LR, Stevens RD, Newgard CB, Farese RV, Jr., de Cabo R, Ulrich S, Akassoglou K, and Verdin E. Suppression of oxidative stress by beta-hydroxybutyrate, an endogenous histone deacetylase inhibitor. Science 339: 211–214, 2013. 155. Song MS, Salmena L, and Pandolfi PP. The functions and regulation of the PTEN tumour suppressor. Nat Rev Mol Cell Biol 13: 283–296, 2012. 156. Sucharov CC, Dockstader K, and McKinsey TA. YY1 protects cardiac myocytes from pathologic hypertrophy by interacting with HDAC5. Mol Biol Cell 19: 4141–4153, 2008. 157. Sun Z, Miller RA, Patel RT, Chen J, Dhir R, Wang H, Zhang D, Graham MJ, Unterman TG, Shulman GI, Sztalryd C, Bennett MJ, Ahima RS, Birnbaum MJ, and Lazar MA. Hepatic Hdac3 promotes gluconeogenesis by repressing lipid synthesis and sequestration. Nat Med 18: 934–942, 2012. 158. Suzuki S, Tanaka T, Poyurovsky MV, Nagano H, Mayama T, Ohkubo S, Lokshin M, Hosokawa H, Nakayama T, Suzuki Y, Sugano S, Sato E, Nagao T, Yokote K, Tatsuno I, and Prives C. Phosphate-activated glutaminase (GLS2), a p53-inducible regulator of glutamine metabolism and reactive oxygen species. Proc Natl Acad Sci U S A 107: 7461–7466, 2010. 158a. Takami Y and Nakayama T. N-terminal region, C-terminal region, nuclear export signal, and deacetylation activity of histone deacetylase-3 are essential for the viability of the DT40 chicken B cell line. J Biol Chem 275: 16191–16201, 2000. 159. Takemori H, Katoh Hashimoto Y, Nakae J, Olson EN, and Okamoto M. Inactivation of HDAC5 by SIK1 in AICARtreated C2C12 myoblasts. Endocr J 56: 121–130, 2009. 160. Tammineni P, Anugula C, Mohammed F, Anjaneyulu M, Larner AC, and Sepuri NB. The import of the transcription factor STAT3 into mitochondria depends on GRIM19, a component of the electron transport chain. J Biol Chem 288: 4723–4732, 2013. 161. Tanoue S, Uto H, Kumamoto R, Arima S, Hashimoto S, Nasu Y, Takami Y, Moriuchi A, Sakiyama T, Oketani M, Ido A, and Tsubouchi H. Liver regeneration after partial hepatectomy in rat is more impaired in a steatotic liver

162. 163.

164.

165. 166.

167.

168.

169. 170.

171.

172.

173.

174.

175.

176.

19

induced by dietary fructose compared to dietary fat. Biochem Biophys Res Commun 407: 163–168, 2011. This reference has been deleted. Taylor SS, Buechler JA, and Yonemoto W. cAMPdependent protein kinase: framework for a diverse family of regulatory enzymes. Annu Rev Biochem 59: 971–1005, 1990. Thangaraju M, Carswell KN, Prasad PD, and Ganapathy V. Colon cancer cells maintain low levels of pyruvate to avoid cell death caused by inhibition of HDAC1/HDAC3. Biochem J 417: 379–389, 2009. van der Heide LP and Smidt MP. Regulation of FoxO activity by CBP/p300-mediated acetylation. Trends Biochem Sci 30: 81–86, 2005. Vervoorts J, Luscher-Firzlaff JM, Rottmann S, Lilischkis R, Walsemann G, Dohmann K, Austen M, and Luscher B. Stimulation of c-MYC transcriptional activity and acetylation by recruitment of the cofactor CBP. EMBO Rep 4: 484–490, 2003. Villagra A, Cheng F, Wang HW, Suarez I, Glozak M, Maurin M, Nguyen D, Wright KL, Atadja PW, Bhalla K, Pinilla-Ibarz J, Seto E, and Sotomayor EM. The histone deacetylase HDAC11 regulates the expression of interleukin 10 and immune tolerance. Nat Immunol 10: 92– 100, 2009. Vogelauer M, Krall AS, McBrian MA, Li JY, and Kurdistani SK. Stimulation of histone deacetylase activity by metabolites of intermediary metabolism. J Biol Chem 287: 32006–32016, 2012. Wagner JM, Hackanson B, Lubbert M, and Jung M. Histone deacetylase (HDAC) inhibitors in recent clinical trials for cancer therapy. Clin Epigenetics 1: 117–136, 2010. Walenta S, Wetterling M, Lehrke M, Schwickert G, Sundfor K, Rofstad EK, and Mueller-Klieser W. High lactate levels predict likelihood of metastases, tumor recurrence, and restricted patient survival in human cervical cancers. Cancer Res 60: 916–921, 2000. Waltregny D, De Leval L, Glenisson W, Ly Tran S, North BJ, Bellahcene A, Weidle U, Verdin E, and Castronovo V. Expression of histone deacetylase 8, a class I histone deacetylase, is restricted to cells showing smooth muscle differentiation in normal human tissues. Am J Pathol 165: 553–564, 2004. Wang F, Chan CH, Chen K, Guan X, Lin HK, and Tong Q. Deacetylation of FOXO3 by SIRT1 or SIRT2 leads to Skp2-mediated FOXO3 ubiquitination and degradation. Oncogene 31: 1546–1557, 2012. Wang GL, Salisbury E, Shi X, Timchenko L, Medrano EE, and Timchenko NA. HDAC1 cooperates with C/ EBPalpha in the inhibition of liver proliferation in old mice. J Biol Chem 283: 26169–26178, 2008. Wang Q, Bailey CG, Ng C, Tiffen J, Thoeng A, Minhas V, Lehman ML, Hendy SC, Buchanan G, Nelson CC, Rasko JE, and Holst J. Androgen receptor and nutrient signaling pathways coordinate the demand for increased amino acid transport during prostate cancer progression. Cancer Res 71: 7525–7536, 2011. Wang Q, Zhang Y, Yang C, Xiong H, Lin Y, Yao J, Li H, Xie L, Zhao W, Yao Y, Ning ZB, Zeng R, Xiong Y, Guan KL, Zhao S, and Zhao GP. Acetylation of metabolic enzymes coordinates carbon source utilization and metabolic flux. Science 327: 1004–1007, 2010. Wardell SE, Ilkayeva OR, Wieman HL, Frigo DE, Rathmell JC, Newgard CB, and McDonnell DP. Glucose me-

20

177. 178.

179.

180.

181. 182. 183. 184.

185.

186.

187.

188. 189. 190. 191. 192.

CHIARADONNA ET AL.

tabolism as a target of histone deacetylase inhibitors. Mol Endocrinol 23: 388–401, 2009. Watabe M and Nakaki T. CK2 as anti-stress factor. Commun Integr Biol 5: 278–280, 2012. Wegrzyn J, Potla R, Chwae YJ, Sepuri NB, Zhang Q, Koeck T, Derecka M, Szczepanek K, Szelag M, Gornicka A, Moh A, Moghaddas S, Chen Q, Bobbili S, Cichy J, Dulak J, Baker DP, Wolfman A, Stuehr D, Hassan MO, Fu XY, Avadhani N, Drake JI, Fawcett P, Lesnefsky EJ, and Larner AC. Function of mitochondrial Stat3 in cellular respiration. Science 323: 793–797, 2009. Wellen KE, Hatzivassiliou G, Sachdeva UM, Bui TV, Cross JR, and Thompson CB. ATP-citrate lyase links cellular metabolism to histone acetylation. Science 324: 1076–1080, 2009. Wise DR, DeBerardinis RJ, Mancuso A, Sayed N, Zhang XY, Pfeiffer HK, Nissim I, Daikhin E, Yudkoff M, McMahon SB, and Thompson CB. Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proc Natl Acad Sci U S A 105: 18782–18787, 2008. Wise DR and Thompson CB. Glutamine addiction: a new therapeutic target in cancer. Trends Biochem Sci 35: 427– 433, 2010. Witt O, Deubzer HE, Milde T, and Oehme I. HDAC family: what are the cancer relevant targets? Cancer Lett 277: 8–21, 2009. Wolfson NA, Ann Pitcairn C, and Fierke CA. HDAC8 substrates: histones and beyond. Biopolymers 99: 112– 126, 2013. Wu S, Kasim V, Kano MR, Tanaka S, Ohba S, Miura Y, Miyata K, Liu X, Matsuhashi A, Chung UI, Yang L, Kataoka K, Nishiyama N, and Miyagishi M. Transcription factor YY1 contributes to tumor growth by stabilizing hypoxia factor HIF-1alpha in a p53-independent manner. Cancer Res 73: 1787–1799, 2013. Wu WS, Vallian S, Seto E, Yang WM, Edmondson D, Roth S, and Chang KS. The growth suppressor PML represses transcription by functionally and physically interacting with histone deacetylases. Mol Cell Biol 21: 2259–2268, 2001. Yang L, Vaitheesvaran B, Hartil K, Robinson AJ, Hoopmann MR, Eng JK, Kurland IJ, and Bruce JE. The fasted/ fed mouse metabolic acetylome: N6-acetylation differences suggest acetylation coordinates organ-specific fuel switching. J Proteome Res 10: 4134–4149, 2011. Yang T, Jian W, Luo Y, Fu X, Noguchi C, Bungert J, Huang S, and Qiu Y. Acetylation of histone deacetylase 1 regulates NuRD corepressor complex activity. J Biol Chem 287: 40279–40291, 2012. Yang W and Lu Z. Regulation and function of pyruvate kinase M2 in cancer. Cancer Lett 339: 153–158, 2013. Yang WM, Tsai SC, Wen YD, Fejer G, and Seto E. Functional domains of histone deacetylase-3. J Biol Chem 277: 9447–9454, 2002. Yang XJ and Seto E. The Rpd3/Hda1 family of lysine deacetylases: from bacteria and yeast to mice and men. Nat Rev Mol Cell Biol 9: 206–218, 2008. Yao YL, Yang WM, and Seto E. Regulation of transcription factor YY1 by acetylation and deacetylation. Mol Cell Biol 21: 5979–5991, 2001. Ye J, Mancuso A, Tong X, Ward PS, Fan J, Rabinowitz JD, and Thompson CB. Pyruvate kinase M2 promotes de novo serine synthesis to sustain mTORC1 activity and cell

193.

194. 195. 196.

197.

198.

199.

200.

proliferation. Proc Natl Acad Sci U S A 109: 6904–6909, 2012. Yoo YG, Kong G, and Lee MO. Metastasis-associated protein 1 enhances stability of hypoxia-inducible factor1alpha protein by recruiting histone deacetylase 1. EMBO J 25: 1231–1241, 2006. Yuan ZL, Guan YJ, Chatterjee D, and Chin YE. Stat3 dimerization regulated by reversible acetylation of a single lysine residue. Science 307: 269–273, 2005. Zentner GE and Henikoff S. Regulation of nucleosome dynamics by histone modifications. Nat Struct Mol Biol 20: 259–266, 2013. Zhang J, Kan S, Huang B, Hao Z, Mak TW, and Zhong Q. Mule determines the apoptotic response to HDAC inhibitors by targeted ubiquitination and destruction of HDAC2. Genes Dev 25: 2610–2618, 2011. Zhang X, Ozawa Y, Lee H, Wen YD, Tan TH, Wadzinski BE, and Seto E. Histone deacetylase 3 (HDAC3) activity is regulated by interaction with protein serine/threonine phosphatase 4. Genes Dev 19: 827–839, 2005. Zhang X, Yuan Z, Zhang Y, Yong S, Salas-Burgos A, Koomen J, Olashaw N, Parsons JT, Yang XJ, Dent SR, Yao TP, Lane WS, and Seto E. HDAC6 modulates cell motility by altering the acetylation level of cortactin. Mol Cell 27: 197–213, 2007. Zhang Y, Kwon S, Yamaguchi T, Cubizolles F, Rousseaux S, Kneissel M, Cao C, Li N, Cheng HL, Chua K, Lombard D, Mizeracki A, Matthias G, Alt FW, Khochbin S, and Matthias P. Mice lacking histone deacetylase 6 have hyperacetylated tubulin but are viable and develop normally. Mol Cell Biol 28: 1688–1701, 2008. Zhou G, Sebhat IK, and Zhang BB. AMPK activators— potential therapeutics for metabolic and other diseases. Acta Physiol (Oxf ) 196: 175–190, 2009.

Address correspondence to: Prof. Ferdinando Chiaradonna Department of Biotechnology and Biosciences University of Milano-Bicocca Piazza Della Scienza 2 Milan 20126 Italy E-mail: [email protected] Date of first submission to ARS Central, January 16, 2014; date of acceptance, February 2, 2014. Abbreviations Used a-KG ¼ a-ketoglutaric acid bOHB ¼ b-hydroxybutyrate AAs ¼ amino acids ACC ¼ acetyl-coenzyme A carboxylase Ac-CoA ¼ acetyl-coenzyme A AcetoAc-CoA ¼ acetoacetyl-CoA Akt, PKB ¼ protein kinase B AMPK ¼ adenosine monophosphate-activated protein kinase But ¼ butyrate CaMK ¼ calcium/calmodulin-dependent kinase CBP ¼ CREB-binding protein CK2 ¼ casein kinase 2

CROSS-TALK BETWEEN HDACS AND METABOLISM IN CANCER

Abbreviations Used Cont. c-Myc ¼ cellular myelocytomatosis viral oncogene homolog CoA ¼ coenzyme A EGFR ¼ epidermal growth factor receptor FoxO ¼ forkhead box class O GA3P ¼ glyceraldehyde 3-phosphate GLUT ¼ glucose transporter GSH ¼ glutathione HAT ¼ histone acetyltransferase HBP ¼ hexosamine biosynthesis pathway HDAC ¼ histone deacetylase HDACi ¼ histone deacetylase inhibitor HIF1a ¼ hypoxia-inducible factor 1a HK ¼ hexokinase HMG-CoA ¼ 3-Hydroxy-3-methylglutarylCoenzyme A KO ¼ knockout LDHA ¼ lactate dehydrogenase A MDM2 ¼ mouse double minute 2 homolog Me-crotonyl-CoA ¼ methylcrotonyl-CoA Me-malonyl-CoA ¼ methylmalonyl-CoA mTOR and mTORC1 ¼ mammalian target of rapamycin and mTOR complex 1 N-Cor ¼ nuclear receptor corepressor NES ¼ nuclear export signal NLS ¼ nuclear localization signal

21

NTs ¼ nucleotides OAA ¼ oxalacetate OXPHOS ¼ oxidative phosphorylation p300 ¼ E1 A-associated protein p300 PCAF ¼ p300/CBP-associated factor PDK ¼ pyruvate dehydrogenase kinase PGC-1a ¼ peroxisome proliferator-activated receptor-c coactivator PKA and PKC ¼ protein kinase A and C PKM2 ¼ pyruvate kinase M2 PPP ¼ pentose phosphate pathway PTEN ¼ phosphatase and tensin homolog deleted in chromosome 10 ROS ¼ reactive oxygen species SAHA ¼ suberoylanilide hydroxamic acid SIK1 ¼ salt-inducible kinase 1 SIRT ¼ class III HDACs SLGT1 ¼ sodium-glucose transport protein 1 SMRT ¼ silencing mediator of retinoid and thyroid receptors STAT3 ¼ signal transducer and activator of transcription 3 TCA ¼ tricarboxylic acid (cycle) TSA ¼ trichostatin A VHL ¼ E3-ubiquitin ligase von Hippel-Lindau YY1 ¼ Yin Yang 1

New Insights into the Connection Between Histone Deacetylases, Cell Metabolism, and Cancer.

Histone deacetylases (HDACs) activity and cell metabolism are considered important targets for cancer therapy, as both are deregulated and associated ...
1MB Sizes 0 Downloads 0 Views