Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Nov 1.
Published in final edited form as: DNA Repair (Amst). 2014 Sep 29;0:64–68. doi: 10.1016/j.dnarep.2014.08.005

The function of nicotinamide phosphoribosyltransferase in the heart

Chiao-Po Hsu 2, Takanobu Yamamoto 1, Shinichi Oka 1, Junichi Sadoshima 1,3
PMCID: PMC4254158  NIHMSID: NIHMS633790  PMID: 25277684

Abstract

In addition to its roles as a coenzyme and an electron transfer molecule, nicotinamide adenine dinucleotide (NAD+) has emerged as a substrate of sirtuins, a family of enzymes that control aging and metabolism. Nicotinamide phosphoribosyltransferase (Nampt), a rate-limiting enzyme in the NAD+ salvage pathway, plays an important role in controlling the level of NAD+ and the activity of Sirt1 in the heart and the cardiomyocytes therein. Nampt protects the heart from ischemia and reperfusion injury by stimulating Sirt1. In this review, we summarize what is currently known regarding the function of Nampt in the heart.

Introduction

Nicotinamide adenine dinucleotide (NAD+) is classically regarded as a coenzyme for energy production and participates in cellular redox reactions as a transfer molecule for electrons. During the past decade, NAD+ was also found to serve as a substrate of sirtuins and activated poly(ADP-ribose) polymerase (PARP).[1, 2] Sirtuins are involved in gene silencing, longevity, senescence, differentiation, and cell survival.[3,4] The nuclear enzyme PARP-1 is involved in DNA repair and maintenance of genomic stability in response to DNA damage.[1]

In mammals, NAD+ is synthesized from amino acids, including tryptophan and aspartic acid, via the de novo pathway,[5] or is taken up from the extracellular space.[6, 7] NAD+ can also be resynthesized from NAD+ metabolites through the salvage pathway.[5] In the salvage pathway of NAD+ biosynthesis, Nampt is identified as a rate-limiting enzyme that converts nicotinamide to nicotinamide mononucleotide (NMN).[8, 9] NAD+ is then synthesized from NMN by NMN adenylyltransferase (Nmnat).[10, 11] Nampt has received attention recently not only because it can increase the intracellular NAD+ level, but also because it is a mammalian functional equivalent of pyrazinamidase/nicotinamidase 1 (PNC1), a master gene that contributes to longevity in yeast. [12]

Nampt

Nampt is a highly conserved 52-kDa protein without a signal sequence,[12, 13] but it exists in most cells and tissues.[12] Nampt is found in both intracellular and extracellular spaces. Intracellular Nampt (iNampt) is primarily involved in NAD+ synthesis, whereas extracellular Nampt (eNampt), secreted from some cell types including neutrophils, adipocytes, cardiomyocytes and mesangial cells, may act as an extracellular NMN synthesis enzyme or as a cytokine/adipokine.[12, 1417] The mechanism through which Nampt is secreted is not well understood. However, Nampt may leak from damaged tissues, which may explain why Nampt serum levels increase in a mouse model of acute myocardial infarction.[18]

The gene encoding Nampt/PBEF was isolated from a human peripheral blood lymphocyte cDNA library in 1994. It was named pre-B-cell colony-enhancing factor (PBEF) because it was thought to be a cytokine that acted on early B-lineage precursor cells.[12] Later, on the basis of sequence similarity between PBEF and nadV, a prokaryotic nicotinamide phosphoribosyltranferase (NAmPRTase) from Haemophilus ducreyi, PBEF was identified as Nampt, an enzyme involved in NAD+ biosynthesis.[9, 19] Recently, visfatin, an adipocytokine proposed to regulate insulin secretion in pancreatic β-cells, was found to be identical to Nampt,[17] although cytokine actions of Nampt require further confirmation. Nampt might be involved in aging and a broad range of disorders, including human immunodeficiency virus infection, septicemia, myocardial failure, atherosclerosis, metabolic disorders, inflammatory diseases, malignancies, and neurodegenerative disorders, through both NAD+ synthesis and cytokine actions.[20]

Nampt controls NAD+ levels in the heart

Downregulation of Nampt significantly decreases NAD+ content in cardiomyocytes at baseline, suggesting that Nampt plays a crucial role in NAD+ synthesis in cardiomyocytes.[21] On the other hand, overexpression of Nampt readily increases NAD+ content, indicating that Nampt is a rate-limiting enzyme for NAD+ synthesis in cardiomyocytes.[21] Interestingly, Nampt is downregulated at both the protein and mRNA levels under stress conditions in the heart, including 24 hours of ischemia, 45 minutes of ischemia followed by 24 hours of reperfusion, and 2 and 4 weeks of pressure overload due to thoracic aortic constriction,[21] which induce compensated and decompensated stages of cardiac hypertrophy, respectively, in mice. These changes are accompanied by decreases in the level of NAD+ in the heart ([22] and our unpublished observation). These results suggest that Nampt is a critical determinant of the NAD+ level in the heart at baseline and in response to stress.[23, 24]

Generation of transgenic mice with cardiac-specific overexpression of Nampt

In order to evaluate the function of Nampt in the heart, transgenic mice with cardiac-specific overexpression of Nampt (Tg-Nampt) were generated. These mice exhibited higher levels of NAD+ and NAD+/NADH than non-transgenic (NTg) control mice at baseline. The baseline cardiac phenotype, including heart weight, chamber size, and left ventricular function, was not significantly different between Tg-Nampt and NTg mice.[21] This observation is different from a recent report by Pillai et al.,[14] in which mice with Nampt overexpression developed cardiac hypertrophy. The reason for the discrepancy remains to be elucidated but the level of Nampt overexpression was greater in Pillai’s report than in ours.

Nampt protects the heart against ischemia/reperfusion (I/R) injury

To examine the effect of Nampt upon myocardial I/R injury, Tg-Nampt mice were subjected to 45 minutes of ischemia followed by 24 hours of reperfusion and the degree of myocardial infarction was evaluated. The size of the myocardial infarct/area at risk was significantly smaller in Tg-Nampt than in NTg mice.[21] Since endogenous Nampt is downregulated by I/R and cardiac-specific overexpression of Nampt restores the NAD+ level,[21] these results suggest that downregulation of endogenous Nampt contributes to myocardial injury in response to I/R and that upregulation of Nampt protects the heart against I/R injury. The number of TUNEL-positive cells in the ischemic border zone was significantly lower in Tg-Nampt mice than in NTg mice, suggesting that Nampt suppresses cardiomyocyte apoptosis.[21] In order to examine whether Nampt has cell-autonomous and protective effects against cell death in cardiomyocytes, we transduced cardiomyocytes with adenovirus harboring Nampt or shRNA-Nampt. Overexpression of Nampt protected cardiomyocytes from cell death induced by either methylmethane sulfonate (MMS), a DNA-alkylating agent known to induce necrotic cell death, [25] or glucose deprivation. On the other hand, downregulation of Nampt significantly increased apoptotic cell death in cardiomyocytes, as evaluated with TUNEL staining and immunoblot analyses of cleaved caspase-3.[21] Furthermore, the increase in cell death induced by downregulation of Nampt was reversed in the presence of adenovirus harboring Bcl-xL, an inhibitor of apoptosis.[21] These results suggest that Nampt directly regulates apoptosis in cardiomyocytes.[21] We also investigated whether knockdown of Nampt causes necrotic cell death. Nampt downregulation alone did not significantly increase the number of necrotic myocytes, as evaluated with propidium iodide staining and in situ ligation with hairpin 2 probes. However, Nampt downregulation significantly increased necrotic cell death in the presence of glucose deprivation or MMS.[21] Taken altogether, these results suggest that although downregulation of Nampt induces apoptotic cell death at baseline, it may facilitate necrotic cell death under stress.

What, then, is the underlying mechanism by which Nampt protects cardiomyocytes from cell death? Upregulation of Nampt increases the cellular NAD+ level and enhances the activity of Sirt1 in mouse fibroblasts.[8, 26] We have shown previously that the size of the myocardial infarct and the number of TUNEL-positive nuclei induced by ischemia/reperfusion are significantly reduced in transgenic mice with cardiac-specific overexpression of Sirt1 compared to NTg mice,[27] whereas cardiac-specific Sirt1 homozygous knockout mice exhibit significantly larger myocardial infarcts and greater numbers of apoptotic cardiomyocytes than NTg mice.[28] Thus, we speculate that Sirt1, an NAD+-dependent enzyme, may play a role in mediating the effect of Nampt in cardiomyocytes.[29]

Although the results described thus far suggest that cardiac-specific overexpression of Nampt protects the heart, the role of eNampt in regulating I/R injury remains unclear. Lim et al. showed that eNampt is capable of reducing myocardial injury when administered at the time of myocardial reperfusion by regulating the PI3K and MEK1/2 pathways and mitochondrial permeability transition pore (mPTP) opening.[30] Xiao et al. showed that pretreatment with eNampt attenuates apoptotic cell death and mitochondrial membrane potential depolarization in H9c2 myocardial cells in response to H2O2 treatment.[31] Although eNampt did not inhibit the death receptor-dependent apoptotic pathway, it suppressed the mitochondria-dependent apoptotic pathway by regulating p53 and Bcl-2 family proteins through AMPK activation.[31] Taken altogether, these results suggest that eNampt may contribute to the protective effect of Nampt against I/R observed in Tg-Nampt. Because Nampt could be released from cardiomyocytes, [14] the relative contribution of iNampt and eNampt to protection against I/R injury in Tg-Nampt remains to be elucidated. Montecucco et al. showed that inhibition of Nampt by FK866, a chemical inhibitor, decreased myocardial infarct size in a mouse model of I/R in vivo by reducing neutrophil-mediated injury through downregulation of CXCL2.[18] This would appear to indicate that Nampt has a detrimental effect upon I/R injury. It is unknown whether eNampt mediates the detrimental effect of Nampt. The NAD+ salvage pathway is regulated by the circadian clock through epigenetic control and chromatin remodeling,[32, 33] and also by AMPK in the heart.[34] Whether NAD+ exhibits circadian changes in the heart, and if so, whether such a mechanism contributes to the well-known circadian variation in the incidence of myocardial infarction in humans requires further investigation.

Nampt protects the heart against I/R injury in part by stimulating autophagy

Since Nampt protects the heart against ischemia, we examined whether Nampt affects autophagy, a cellular mechanism of degradation to preserve cellular ATP levels and/or to achieve quality control of proteins and organelles.[35] Nampt knockdown (Ad-shRNA-Nampt) increased autophagosomes in cardiomyocytes, as assessed by counting of GFP-LC3 puncta, which allow visualization of autophagosome formation, and by LC3-II immunoblot analyses.[21] Importantly, however, the accumulation of GFP-LC3 puncta observed in the presence of Nampt knockdown was not enhanced in the presence of chloroquine,[21] an inhibitor of autophagosome-lysosome fusion. p62/SQSTM1, a protein known to be degraded by autophagy, [3638] was accumulated in the presence of Nampt knockdown, whereas the mRNA level of p62 was not significantly affected.[21] These results suggest that downregulation of Nampt inhibits autophagic flux in cardiomyocytes. We speculate that NAD+ synthesis by Nampt plays an important role in mediating either fusion between autophagosomes and lysosomes or lysosomal degradation of cargo proteins.

Supplementation with exogenous NAD+ significantly reversed the accumulation of autophagosomes in the presence of Nampt knockdown,[21] suggesting that the effect of Nampt downregulation is mediated through decreases in NAD+. Since Sirt1 stimulates autophagy in HeLa and MEF cells through deacetylation of autophagy-related proteins,[39] we examined the effect of Sirt1 knockdown upon autophagy in cardiomyocytes. Downregulation of Sirt1 caused accumulation of LC3-II and p62, thereby mimicking the effect of Nampt downregulation.[21] Furthermore, double knockdown of Nampt and Sirt1 did not show an additive effect upon accumulation of LC3-II and p62,[21] suggesting that Nampt and Sirt1 affect autophagy via a common mechanism.

Since downregulation of Nampt inhibits autophagic flux, we hypothesized that overexpression of Nampt during ischemic stress may protect the heart by stimulating autophagic flux. The size of the myocardial infarct/area at risk after 2 hours of ischemia was significantly smaller in Tg-Nampt than in NTg.[21] Furthermore, the level of autophagic flux was greater in Tg-Nampt than in NTg, as evidenced by less accumulation of p62 in Tg-Nampt.[21] Taken altogether, these results are consistent with the notion that the downregulation of endogenous Nampt observed in the heart during ischemia is detrimental partly due to inhibition of autophagic flux. We speculate that Nampt-induced increases in autophagic flux most likely occur through stimulation of Sirt1 in cardiomyocytes, since Sirt1 has essentially the same effect as Nampt upon autophagy. Furthermore, the close similarity of the in vivo phenotypes of Tg-Nampt and Tg-Sirt1 in response to stress [21, 27] suggests that there may be a close functional coupling between Nampt and Sirt1 (sirtuins) in the heart even if they may not have direct physical interaction.

Nampt inhibition by the specific chemical inhibitor FK866 also increases autophagosome formation, as evaluated with GFP-LC3 puncta, in SH-SY5Y neuroblastoma cells. However, since autophagosome accumulation can be observed when autophagic flux is blocked, whether FK866 stimulates autophagic flux remains to be clarified.[6] Recently, Cea et al. showed that Nampt plays an essential role in maintaining both cell viability and intracellular NAD+ stores in multiple myeloma cell lines, and that FK866 triggers autophagic, but not apoptotic, cell death.[40] Again, it would be critical to test whether suppression of Nampt by FK866 really stimulates autophagic flux in cancer cells. Nampt is upregulated in cancer cells, which rely on NAD+ to support their rapid cell proliferation.[4143] Interestingly, Nampt promotes survival of neurons, terminally differentiated cells, during cerebral ischemia. This occurs through induction of autophagy via regulation of the TSC2-mTOR-S6K1 pathway in a Sirt1-dependent manner.[44] Thus, Nampt may have distinct effects upon autophagy in terminally differentiated cells and cancer cells.

The effect of Nampt upon cardiac hypertrophy

Pillai et al. showed that agonist-induced cardiac hypertrophy is associated with loss of intracellular levels of NAD+. Exogenous addition of NAD+ was capable of maintaining intracellular levels of NAD+ and blocking the agonist-induced cardiac hypertrophy in vitro and in vivo through activation of Sirt3.[22] Whether the suppression of hypertrophy is due to the direct effect of NAD+ upon the hypertrophy machinery or is secondary to the improvement of cardiac function due to normalization of NAD+ content remains to be elucidated. Cai et al. showed that maintenance of the intracellular NAD+ level by overexpression of Nmnat2, a form of Nmnat located in the cytoplasm, blocks angiotensin II-induced cardiac hypertrophy in neonatal rat cardiomyocytes.[45] The authors did not evaluate the tissue level of NAD+ in these transgenic mice. However, these results suggest that Nampt is protective against pressure-overload-induced cardiac hypertrophy. Recently, Karamanlidis et al. showed that complex I deficiency caused by genetic deletion of Ndufs4 causes accumulation of NADH and a decrease in the NAD+/NADH ratio, which in turn inhibits Sirt3 and induces hyperacetylation of mitochondrial proteins and mPTP opening. These changes were alleviated by normalization of the NAD+/NADH ratio. The complex I deficiency rendered mice highly susceptible to additional stresses and accelerated the development of heart failure after pressure overload or repeated pregnancy. [46] Thus, maintaining the intracellular NAD+ level may suppress cardiac hypertrophy and failure. Supplementation of NAD+ or enzymes to produce NAD+ may be considered for treatment of cardiac hypertrophy and heart failure in the future. Although external application of NAD+ has been used to increase the level of NAD+ and stimulate NAD+-dependent enzymes in the heart or cardiomyocytes, [21,22] the underlying mechanism of uptake and its efficacy are not well understood.

In contrast, using cardiac-specific Nampt overexpression mice, Pillai et al. demonstrated that cardiomyocytes are capable of secreting Nampt during stress and that Nampt is a facilitator of stress-induced cardiac hypertrophy.[14] It should be noted that cardiac hypertrophy has never been observed in the cardiac-specific Nampt overexpression mice we generated.[21] Nampt stimulates NF-κB, whereas it also inhibits tumor necrosis factor-alpha (TNF-alpha)-induced NF-κB activity.[47] eNampt may protect cells from ER-stress-induced apoptosis by activating an IL-6/STAT3 signaling pathway through a non-enzymatic mechanism.[48] Thus, Nampt appears to have pleiotropic actions and its function may be context-dependent in the heart.[24] Although whether Nampt is secreted from cardiomyocytes remains to be clarified, eNampt may act on non-myocytes, which in turn affect growth and death of cardiomyocytes through paracrine mechanisms. [22]

The effect of Nampt upon cardiac metabolism

Until now, there has been no specific report regarding how Nampt affects cardiac metabolism in response to metabolic perturbation, such as high-fat feeding or nutrient deprivation. Canto et al. showed that dietary supplementation with nicotinamide riboside (NR), a vitamin B3 and NAD+ precursor, in mammalian cells and mouse tissues increases NAD+ levels and activates Sirt1 and Sirt3, culminating in enhanced oxidative metabolism and protection against high-fat-diet-induced metabolic abnormalities.[49] Khan et al. also found that NR increases mitochondrial biogenesis in skeletal muscle and ameliorates mitochondrial myopathy.[50] Yoshino et al. also demonstrated that nicotinamide mononucleotide (NMN) improves glucose intolerance and lipid profiles in high-fat-diet- and age-induced type 2 diabetic mice, partly through Sirt1 activation.[51] These results show that NAD+ and its precursors may have significant effects on metabolism in response to nutrient stress, and that their effects may be attributable to sirtuins in the heart as well. Although Nampt would be expected to exert similar effects on cardiac metabolism, more experiments are needed to prove this hypothesis.

Conclusions

Nampt critically regulates NAD+ content, thereby playing an essential role in mediating cell survival by inhibiting apoptosis and stimulating autophagic flux in cardiomyocytes. Preventing downregulation of Nampt inhibits myocardial injury in response to myocardial ischemia and reperfusion as well as other stress conditions in the heart (Table). Although supplementation of NAD+ through upregulation of Nampt is generally protective in the heart, the role of Nampt in the heart during stress requires further clarification because eNampt may have cytokine actions that may either positively or negatively affect the survival of cardiomyocytes.

Table.

The cardiac effects of Nampt/NAD+ modulation in different models of heart dysfunction

Modulation of
  Nampt/NAD+
Stress model Cardiac effects Reference
Nampt
heterozygous (+/−)
knockout
iNampt
overexpression
(Transgenic mice)
hypertrophic (isoproterenol
and angiotensin II) stimuli

Baseline mice without
stimulus
Protection against
hypertrophy


Development cardiac
hypertrophy and fibrosis at 6
months of age
[14]
Administration of
FK866 (inhibitor
of Nampt)
Myocardial
ischemia/reperfusion
model (in vivo and ex vivo)
in mouse
Reduction of myocardial
injury only in vivo, but not ex
vivo (Reduce
neutrophil-mediated injury)
[18]
iNampt
overexpression
(Transgenic mice)
Myocardial ischemia and
ischemia/reperfusion
model
Reduction of myocardial
injury
Stimulation of autophagy
[21]
Exogenous
addition of NAD+
hypertrophic (angiotensin
II) stimuli
Inhibition of cardiac
hypertrophy through
activation of Sirt3
[22]
Administration of
Nampt (eNampt) at
the time of
reperfusion
Myocardial
ischemia/reperfusion
model
Reduction of myocardial
injury
[30]
Administration of
Nampt (eNampt)
Before H2O2 treatment in
H9c2 cardiomyocytes
Suppression of myocardial
apoptosis
[31]
Overexpression of
Nmnat2
hypertrophic (angiotensin
II) stimuli
Inhibition of cardiac
hypertrophy in neonatal rat
cardiomyocytes
[45]
Decrease the
NAD+/NADH
ratio by
cardiac-specific
deletion of Ndufs4
Pressure overload
(Transverse aortic
constriction) or repeated
pregnancy.
Acceleration of heart failure [46]

Acknowledgements

The authors wish to thank Daniela Zablocki, Christopher D. Brady and Dominic P. Del Re for critical reading of the manuscript and suggestions. This work was supported in part by U.S. Public Health Service Grants HL102738, HL67724, HL69020, HL91469, AG23039, and AG27211. This work was also supported by the Fondation Leducq Transatlantic Networks of Excellence.

Footnotes

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Virag L, Szabo C. The therapeutic potential of poly(ADP-ribose) polymerase inhibitors. Pharmacol Rev. 2002;54:375–429. doi: 10.1124/pr.54.3.375. [DOI] [PubMed] [Google Scholar]
  • 2.Imai S, Armstrong CM, Kaeberlein M, Guarente L. Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature. 2000;403:795–800. doi: 10.1038/35001622. [DOI] [PubMed] [Google Scholar]
  • 3.Guarente L. Sir2 links chromatin silencing, metabolism, and aging. Genes Dev. 2000;14:1021–1026. [PubMed] [Google Scholar]
  • 4.Ha CW, Huh WK. The implication of Sir2 in replicative aging and senescence in Saccharomyces cerevisiae. Aging (Albany NY) 2011;3:319–324. doi: 10.18632/aging.100299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Revollo JR, Grimm AA, Imai S. The regulation of nicotinamide adenine dinucleotide biosynthesis by Nampt/PBEF/visfatin in mammals. Curr Opin Gastroenterol. 2007;23:164–170. doi: 10.1097/MOG.0b013e32801b3c8f. [DOI] [PubMed] [Google Scholar]
  • 6.Billington RA, Genazzani AA, Travelli C, Condorelli F. NAD depletion by FK866 induces autophagy. Autophagy. 2008;4:385–387. doi: 10.4161/auto.5635. [DOI] [PubMed] [Google Scholar]
  • 7.Ying W, Garnier P, Swanson RA. NAD+ repletion prevents PARP-1-induced glycolytic blockade and cell death in cultured mouse astrocytes. Biochem Biophys Res Commun. 2003;308:809–813. doi: 10.1016/s0006-291x(03)01483-9. [DOI] [PubMed] [Google Scholar]
  • 8.Revollo JR, Grimm AA, Imai S. The NAD biosynthesis pathway mediated by nicotinamide phosphoribosyltransferase regulates Sir2 activity in mammalian cells. J Biol Chem. 2004;279:50754–50763. doi: 10.1074/jbc.M408388200. [DOI] [PubMed] [Google Scholar]
  • 9.Rongvaux A, Shea RJ, Mulks MH, Gigot D, Urbain J, Leo O, Andris F. Pre-B-cell colony-enhancing factor, whose expression is up-regulated in activated lymphocytes, is a nicotinamide phosphoribosyltransferase, a cytosolic enzyme involved in NAD biosynthesis. Eur J Immunol. 2002;32:3225–3234. doi: 10.1002/1521-4141(200211)32:11<3225::AID-IMMU3225>3.0.CO;2-L. [DOI] [PubMed] [Google Scholar]
  • 10.Emanuelli M, Carnevali F, Saccucci F, Pierella F, Amici A, Raffaelli N, Magni G. Molecular cloning, chromosomal localization, tissue mRNA levels, bacterial expression, and enzymatic properties of human NMN adenylyltransferase. J Biol Chem. 2001;276:406–412. doi: 10.1074/jbc.M008700200. [DOI] [PubMed] [Google Scholar]
  • 11.Schweiger M, Hennig K, Lerner F, Niere M, Hirsch-Kauffmann M, Specht T, Weise C, Oei SL, Ziegler M. Characterization of recombinant human nicotinamide mononucleotide adenylyl transferase (NMNAT), a nuclear enzyme essential for NAD synthesis. FEBS Lett. 2001;492:95–100. doi: 10.1016/s0014-5793(01)02180-9. [DOI] [PubMed] [Google Scholar]
  • 12.Samal B, Sun Y, Stearns G, Xie C, Suggs S, McNiece I. Cloning and characterization of the cDNA encoding a novel human pre-B-cell colony-enhancing factor. Mol Cell Biol. 1994;14:1431–1437. doi: 10.1128/mcb.14.2.1431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Tanaka M, Nozaki M, Fukuhara A, Segawa K, Aoki N, Matsuda M, Komuro R, Shimomura I. Visfatin is released from 3T3-L1 adipocytes via a non-classical pathway. Biochem Biophys Res Commun. 2007;359:194–201. doi: 10.1016/j.bbrc.2007.05.096. [DOI] [PubMed] [Google Scholar]
  • 14.Pillai VB, Sundaresan NR, Kim G, Samant S, Moreno-Vinasco L, Garcia JG, Gupta MP. Nampt secreted from cardiomyocytes promotes development of cardiac hypertrophy and adverse ventricular remodeling. Am J Physiol Heart Circ Physiol. 2013;304:H415–H426. doi: 10.1152/ajpheart.00468.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Fukuhara A, Matsuda M, Nishizawa M, Segawa K, Tanaka M, Kishimoto K, Matsuki Y, Murakami M, Ichisaka T, Murakami H, Watanabe E, Takagi T, Akiyoshi M, Ohtsubo T, Kihara S, Yamashita S, Makishima M, Funahashi T, Yamanaka S, Hiramatsu R, Matsuzawa Y, Shimomura I. Visfatin: a protein secreted by visceral fat that mimics the effects of insulin. Science. 2005;307:426–430. doi: 10.1126/science.1097243. [DOI] [PubMed] [Google Scholar]
  • 16.Song HK, Lee MH, Kim BK, Park YG, Ko GJ, Kang YS, Han JY, Han SY, Han KH, Kim HK, Cha DR. Visfatin: a new player in mesangial cell physiology and diabetic nephropathy. Am J Physiol Renal Physiol. 2008;295:F1485–F1494. doi: 10.1152/ajprenal.90231.2008. [DOI] [PubMed] [Google Scholar]
  • 17.Revollo JR, Korner A, Mills KF, Satoh A, Wang T, Garten A, Dasgupta B, Sasaki Y, Wolberger C, Townsend RR, Milbrandt J, Kiess W, Imai S. Nampt/PBEF/Visfatin regulates insulin secretion in beta cells as a systemic NAD biosynthetic enzyme. Cell Metab. 2007;6:363–375. doi: 10.1016/j.cmet.2007.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Montecucco F, Bauer I, Braunersreuther V, Bruzzone S, Akhmedov A, Luscher TF, Speer T, Poggi A, Mannino E, Pelli G, Galan K, Bertolotto M, Lenglet S, Garuti A, Montessuit C, Lerch R, Pellieux C, Vuilleumier N, Dallegri F, Mage J, Sebastian C, Mostoslavsky R, Gayet-Ageron A, Patrone F, Mach F, Nencioni A. Inhibition of nicotinamide phosphoribosyltransferase reduces neutrophil-mediated injury in myocardial infarction. Antioxid Redox Signal. 2013;18:630–641. doi: 10.1089/ars.2011.4487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Martin PR, Shea RJ, Mulks MH. Identification of a plasmid-encoded gene from Haemophilus ducreyi which confers NAD independence. J Bacteriol. 2001;183:1168–1174. doi: 10.1128/JB.183.4.1168-1174.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Dahl TB, Holm S, Aukrust P, Halvorsen B. Visfatin/NAMPT: a multifaceted molecule with diverse roles in physiology and pathophysiology. Annu Rev Nutr. 2012;32:229–243. doi: 10.1146/annurev-nutr-071811-150746. [DOI] [PubMed] [Google Scholar]
  • 21.Hsu CP, Oka S, Shao D, Hariharan N, Sadoshima J. Nicotinamide phosphoribosyltransferase regulates cell survival through NAD+ synthesis in cardiac myocytes. Circ Res. 2009;105:481–491. doi: 10.1161/CIRCRESAHA.109.203703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Pillai VB, Sundaresan NR, Kim G, Gupta M, Rajamohan SB, Pillai JB, Samant S, Ravindra PV, Isbatan A, Gupta MP. Exogenous NAD blocks cardiac hypertrophic response via activation of the SIRT3-LKB1-AMP-activated kinase pathway. J Biol Chem. 2010;285:3133–3144. doi: 10.1074/jbc.M109.077271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Nakamura M, Bhatnagar A, Sadoshima J. Overview of pyridine nucleotides review series. Circ Res. 2012;111:604–610. doi: 10.1161/CIRCRESAHA.111.247924. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Oka S, Hsu CP, Sadoshima J. Regulation of cell survival and death by pyridine nucleotides. Circ Res. 2012;111:611–627. doi: 10.1161/CIRCRESAHA.111.247932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Horton JK, Stefanick DF, Wilson SH. Involvement of poly(ADP-ribose) polymerase activity in regulating Chk1-dependent apoptotic cell death. DNA Repair (Amst) 2005;4:1111–1120. doi: 10.1016/j.dnarep.2005.05.011. [DOI] [PubMed] [Google Scholar]
  • 26.Anderson RM, Bitterman KJ, Wood JG, Medvedik O, Sinclair DA. Nicotinamide and PNC1 govern lifespan extension by calorie restriction in Saccharomyces cerevisiae. Nature. 2003;423:181–185. doi: 10.1038/nature01578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Hsu CP, Zhai P, Yamamoto T, Maejima Y, Matsushima S, Hariharan N, Shao D, Takagi H, Oka S, Sadoshima J. Silent information regulator 1 protects the heart from ischemia/reperfusion. Circulation. 2010;122:2170–2182. doi: 10.1161/CIRCULATIONAHA.110.958033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Alcendor RR, Kirshenbaum LA, Imai S, Sadoshima J. Sir2α a longevity factor and a class III histone decetylase, is an essential endogenous inhibitor of apoptosis in cardiac myocytes. Circ Res. 2004;95:971–980. doi: 10.1161/01.RES.0000147557.75257.ff. [DOI] [PubMed] [Google Scholar]
  • 29.Yamamoto T, Sadoshima J. Protection of the heart against ischemia/reperfusion by silent information regulator 1. Trends Cardiovasc Med. 2011;21:27–32. doi: 10.1016/j.tcm.2012.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Lim SY, Davidson SM, Paramanathan AJ, Smith CC, Yellon DM, Hausenloy DJ. The novel adipocytokine visfatin exerts direct cardioprotective effects. J Cell Mol Med. 2008;12:1395–1403. doi: 10.1111/j.1582-4934.2008.00332.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Xiao J, Sun B, Li M, Wu Y, Sun XB. A novel adipocytokine visfatin protects against H(2)O(2) -induced myocardial apoptosis: a missing link between obesity and cardiovascular disease. J Cell Physiol. 2013;228:495–501. doi: 10.1002/jcp.24257. [DOI] [PubMed] [Google Scholar]
  • 32.Ramsey KM, Yoshino J, Brace CS, Abrassart D, Kobayashi Y, Marcheva B, Hong HK, Chong JL, Buhr ED, Lee C, Takahashi JS, Imai S, Bass J. Circadian clock feedback cycle through NAMPT-mediated NAD+ biosynthesis. Science. 2009;324:651–654. doi: 10.1126/science.1171641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Nakahata Y, Sahar S, Astarita G, Kaluzova M, Sassone-Corsi P. Circadian control of the NAD+ salvage pathway by CLOCK-SIRT1. Science. 2009;324:654–657. doi: 10.1126/science.1170803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Um JH, Pendergast JS, Springer DA, Foretz M, Viollet B, Brown A, Kim MK, Yamazaki S, Chung JH. AMPK regulates circadian rhythms in a tissue- and isoform-specific manner. PLoS ONE. 2011;6:e18450. doi: 10.1371/journal.pone.0018450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Mizushima N, Klionsky DJ. Protein turnover via autophagy: implications for metabolism. Annu Rev Nutr. 2007;27:19–40. doi: 10.1146/annurev.nutr.27.061406.093749. [DOI] [PubMed] [Google Scholar]
  • 36.Bjorkoy G, Lamark T, Johansen T. p62/SQSTM1: a missing link between protein aggregates and the autophagy machinery. Autophagy. 2006;2:138–139. doi: 10.4161/auto.2.2.2405. [DOI] [PubMed] [Google Scholar]
  • 37.Nakai A, Yamaguchi O, Takeda T, Higuchi Y, Hikoso S, Taniike M, Omiya S, Mizote I, Matsumura Y, Asahi M, Nishida K, Hori M, Mizushima N, Otsu K. The role of autophagy in cardiomyocytes in the basal state and in response to hemodynamic stress. Nat Med. 2007;13:619–624. doi: 10.1038/nm1574. [DOI] [PubMed] [Google Scholar]
  • 38.Wang QJ, Ding Y, Kohtz DS, Mizushima N, Cristea IM, Rout MP, Chait BT, Zhong Y, Heintz N, Yue Z. Induction of autophagy in axonal dystrophy and degeneration. J Neurosci. 2006;26:8057–8068. doi: 10.1523/JNEUROSCI.2261-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Lee IH, Cao L, Mostoslavsky R, Lombard DB, Liu J, Bruns NE, Tsokos M, Alt FW, Finkel T. A role for the NAD-dependent deacetylase Sirt1 in the regulation of autophagy. Proc Natl Acad Sci U S A. 2008;105:3374–3379. doi: 10.1073/pnas.0712145105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Cea M, Cagnetta A, Fulciniti M, Tai YT, Hideshima T, Chauhan D, Roccaro A, Sacco A, Calimeri T, Cottini F, Jakubikova J, Kong SY, Patrone F, Nencioni A, Gobbi M, Richardson P, Munshi N, Anderson KC. Targeting NAD+ salvage pathway induces autophagy in multiple myeloma cells via mTORC1 and extracellular signal-regulated kinase (ERK1/2) inhibition. Blood. 2012;120:3519–3529. doi: 10.1182/blood-2012-03-416776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Belenky P, Bogan KL, Brenner C. NAD+ metabolism in health and disease. Trends Biochem Sci. 2007;32:12–19. doi: 10.1016/j.tibs.2006.11.006. [DOI] [PubMed] [Google Scholar]
  • 42.Van Beijnum JR, Moerkerk PT, Gerbers AJ, De Bruine AP, Arends JW, Hoogenboom HR, Hufton SE. Target validation for genomics using peptide-specific phage antibodies: a study of five gene products overexpressed in colorectal cancer. Int J Cancer. 2002;101:118–127. doi: 10.1002/ijc.10584. [DOI] [PubMed] [Google Scholar]
  • 43.Hufton SE, Moerkerk PT, Brandwijk R, de Bruine AP, Arends JW, Hoogenboom HR. A profile of differentially expressed genes in primary colorectal cancer using suppression subtractive hybridization. FEBS Lett. 1999;463:77–82. doi: 10.1016/s0014-5793(99)01578-1. [DOI] [PubMed] [Google Scholar]
  • 44.Wang P, Guan YF, Du H, Zhai QW, Su DF, Miao CY. Induction of autophagy contributes to the neuroprotection of nicotinamide phosphoribosyltransferase in cerebral ischemia. Autophagy. 2012;8:77–87. doi: 10.4161/auto.8.1.18274. [DOI] [PubMed] [Google Scholar]
  • 45.Cai Y, Yu SS, Chen SR, Pi RB, Gao S, Li H, Ye JT, Liu PQ. Nmnat2 protects cardiomyocytes from hypertrophy via activation of SIRT6. FEBS Lett. 2012;586:866–874. doi: 10.1016/j.febslet.2012.02.014. [DOI] [PubMed] [Google Scholar]
  • 46.Karamanlidis G, Lee CF, Garcia-Menendez L, Kolwicz SC, Jr, Suthammarak W, Gong G, Sedensky MM, Morgan PG, Wang W, Tian R. Mitochondrial complex I deficiency increases protein acetylation and accelerates heart failure. Cell Metab. 2013;18:239–250. doi: 10.1016/j.cmet.2013.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Adya R, Tan BK, Chen J, Randeva HS. Nuclear factor-kappaB induction by visfatin in human vascular endothelial cells: its role in MMP-2/9 production and activation. Diabetes Care. 2008;31:758–760. doi: 10.2337/dc07-1544. [DOI] [PubMed] [Google Scholar]
  • 48.Li Y, Zhang Y, Dorweiler B, Cui D, Wang T, Woo CW, Brunkan CS, Wolberger C, Imai S, Tabas I. Extracellular Nampt promotes macrophage survival via a nonenzymatic interleukin-6/STAT3 signaling mechanism. J Biol Chem. 2008;283:34833–34843. doi: 10.1074/jbc.M805866200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Canto C, Houtkooper RH, Pirinen E, Youn DY, Oosterveer MH, Cen Y, Fernandez-Marcos PJ, Yamamoto H, Andreux PA, Cettour-Rose P, Gademann K, Rinsch C, Schoonjans K, Sauve AA, Auwerx J. The NAD(+) precursor nicotinamide riboside enhances oxidative metabolism and protects against high-fat diet-induced obesity. Cell Metab. 2012;15:838–847. doi: 10.1016/j.cmet.2012.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Khan NA, Auranen M, Paetau I, Pirinen E, Euro L, Forsstrom S, Pasila L, Velagapudi V, Carroll CJ, Auwerx J, Suomalainen A. Effective treatment of mitochondrial myopathy by nicotinamide riboside, a vitamin B3. EMBO Mol Med. 2014 doi: 10.1002/emmm.201403943. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Yoshino J, Mills KF, Yoon MJ, Imai S. Nicotinamide mononucleotide, a key NAD(+) intermediate, treats the pathophysiology of diet- and age-induced diabetes in mice. Cell Metab. 2011;14:528–536. doi: 10.1016/j.cmet.2011.08.014. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES