You are on page 1of 27

https://doi.org/10.

1038/s41580-020-00313-x

Supplementary information

NAD+ metabolism and its roles in cellular


processes during ageing
In the format provided by the
authors and unedited
Supplementary Box 1. NAD+ subcellular localization

Nicotinamide adenine dinucleotide (NAD+) functions both as a redox cofactor and as a substrate
for several families of enzymes and its intracellular localization is highly compartmentalized in the
cytoplasm, mitochondria and nucleus, which represent the main subcellular pools. Alternative
NAD+ compartments such as endoplasmic reticulum and peroxisome have been described but
remain largely uncharacterized 1. Additionally, there is evidence that intracellular NAD+
localization and concentration is not uniform, with the highest concentration of NAD+ being in the
mitochondria, and a lesser amount in the nuclear-cytosol space 2. Thus, a major question in the
field has been how these subcellular NAD+ pools are integrated and how they are differentially
regulated and involved in distinct NAD+-dependent processes. Consistent with each NAD+ pool
being regulated independently, the enzymes involved in the biosynthesis or degradation of NAD+
are highly compartmentalized as well.
Since most studies utilize whole-cell and tissue lysates to measure NAD+ via liquid
chromatography-Mass Spectrometry (LC-MS) and enzymatic assays, it has been extremely
challenging to study NAD+ levels in subcellular compartments and organelles. However, a novel
genetically encoded fluorescent biosensor, which was recently developed, has allowed for the
detection of NAD+ levels in multiple cellular compartments, and a greater understanding how the
subcellular NAD+ pools are maintained and regulated 2 . Using this novel tool, it was validated
that the cytosolic and nuclear NAD+ pools are considered to be exchangeable via diffusion through
the nuclear pore and are characterized by a free NAD+ concentration of around 100uM 2. The
largest NAD+ pool is in the mitochondria with concentrations around 250uM 2. In the mitochondria,
NAD+ was thought to be sequestered due to the impermeability of the inner mitochondrial
membrane and the lack of mitochondrial transporters. After the discovery of mitochondrial NAD+
transporters in plants, yeast, and bacteria, the mammalian NAD+ transporter SLC25A51 has been
recently identified 3,4. The identification of the mitochondrial NAD+ mammalian transporter
confirms previous studies suggesting that intact NAD+ can be transported into mitochondria from
the cytosolic pool 2,5. The development of second-generation NAD+ biosensors in tandem with
isotopic labeling and other novel technologies to study NAD+ levels rapidly and accurately 6,7 will
provide the NAD+ field with valuable tools to investigate the complex cellular NAD+ dynamics in a
more precise and refined manner. These will lead to even more breakthroughs and exciting
discoveries to reveal the major biosynthetic pathways, transporters, and enzymes that control
NAD+ production vs consumption rates in different cells, tissues and disease states.
Intracellular NAD+ can be released into the extracellular space upon cell lysis 8,9 or can
be actively released from various cell types, such as neuronal cells upon stimulation 10,11 or
prostate cancer cells 12. Thus, NAD+ can be found at submicromolar concentrations in an
extracellular pool which is largely uncharacterized 13,14. This extracellular NAD+ is mostly thought
to be degraded, and these by-products are imported inside the cells and converted back to NAD+
(see section “Cellular NAD+ metabolism” and Figure 1). However, the dynamic exchange of
intracellular pools of NAD+ to the extracellular space suggests a role for NAD+ as a potential
endocrine or paracrine signaling molecule. In support of NAD+ acting as a paracrine signal, a
direct exchange through the connexin 43 hemichannels on the plasma membrane has been
shown15. However, the biological function of extracellular NAD+ or cell to cell NAD+ exchange
between the intracellular and extracellular pools remains unclear. Although, it has been proposed
that extracellular NAD+ may also act as a danger signal, activating purinergic receptors similar to
ATP, to activate both innate and adaptive immune cells to initiate inflammatory responses 16,17.
Thus, more research is needed to understand what role extra cellular and subcellular tracking of
NAD+ plays in cell biology.
Supplementary Box 2. Cellular functions of NAD+ and their links to ageing

Metabolic reactions
NAD+ exists in two forms including in an oxidized state known as NAD+, and its reduced form
NADH. NAD+, as well as its phosphorylated version NADP+, is able to accept two electrons in the
form of a hydride ion, making this molecule a critical redox coenzyme essential for redox
homeostasis and multiple cellular functions. One of the most important roles of NAD+ in our cells,
if not the most important, is the role of NAD+ in the catabolism of nutrients in metabolic reactions
such as glycolysis, TCA cycle, glutaminolysis, and β-oxidation. Conversely, NADP+, and more
specifically its reduced version NADPH is primarily used for anabolic processes such as lipid and
cholesterol synthesis, as well as in ROS production and maintenance of antioxidant defense
mechanisms such as reducing glutathione.
During aerobic glycolysis high energy hydride ions are harvested from glucose. NAD+ is
critical for the sixth step of glycolysis, conversion of glyceraldehyde 3-phosphate to D-glycerate
1,3-bisphosphate by the enzyme glyceraldehyde 3-phosphate dehydrogenase (GAPDH), which
utilizes NAD+ as an electron acceptor forming NADH. However, during periods of high rates of
glucose metabolism, such that occur during the Warburg effect, NAD+/NADH ratios can become
too low and NADH must be reoxidized to sustain glycolysis. Thus, our cells have adapted multiple
mechanisms to restore NAD+/NADH ratios to regulate metabolic flux of carbons. For example,
electrons from cytosolic NADH (which is impermeable to the mitochondrial membrane) can be
transferred to NAD+ molecules or the electron transport chain in the mitochondria via the malate-
aspartate shuttle or the glycerol-3-phosphate shuttle, respectively, where they will be used to
make ATP via oxidative phosphorylation. Additionally, NAD+ can be restored in the cytosol by the
diversion of pyruvate to lactate by lactate dehydrogenase (LDH) which converts NADH to NAD+.
Another mechanism recently shown to regulate NAD+ redox balance is via the ability of nuclear
NAD+ levels to regulate the expression of mitochondrial ox-phos genes, via a SIRT1-HIF1a
dependent mechanism, affecting the NAD+/NADH redox state in the mitochondria 18.
This highly orchestrated maintenance of the NAD+/NADH ratio can effectively control the
rate of glycolysis, as well as other metabolic reactions that utilize dehydrogenase enzymes such
as glutamate dehydrogenase which is necessary for glutamine catabolism, the TCA cycle
dehydrogenase enzymes (isocitrate dehydrogenase, a-ketoglutarate dehydrogenase, and malate
dehydrogenase) and in β-oxidation (3-hydroxy-acyl-CoA dehydrogenase). It is important to note
that when the NAD+/NADH ratio becomes too high these reactions can also occur in the opposite
reaction, for example LDH can catalyze the conversion of lactate to pyruvate, converting NAD+ to
NADH in the process. Interestingly, the redox state of NAD may also become dysregulated during
aging (see Redox Reactions supplementary section), however what impact this may have on
metabolic reactions is still unclear and remains an interesting area of future investigation.
Finally, the ability of NAD+ to act as a hydride acceptor and NADH to act as an electron
donor, makes NAD+ an energy reservoir and an efficient carrier of energy from one organelle to
another and one chemical reaction to another 19. Ultimately the potential energy stored in NADH
can be cashed in by transferring the electrons to the electron transport chain in the mitochondria
to make ATP using oxidative phosphorylation, with oxygen as the final electron acceptor.

Redox reactions
During ageing, absolute NAD pools and the NAD+/NADH ratio have been shown to decline in
multiple tissues and across species during the natural ageing process and in progeria mouse
models 20–23. Although increased consumption of NAD+ by NADase enzymes such as CD38 and
PARP1, two of the major NADase responsible for degrading NAD+ during ageing, may explain
why total NAD pools decline, it is not clear why NAD+/NADH ratios are affected during ageing.
Enhanced activity of PARP and CD38 have been shown to significantly lower the NAD+/NADH
ratio compared to WT mice, suggesting enhanced NADase activity is sufficient to not only
decrease the absolute NAD+ levels but also affect the NAD redox state 21,24. In addition to
increased NADase activity, another potential mechanisms is that in ageing mice the decline in
nuclear NAD+ levels inhibits SIRT1 activity and this leads to reduced gene expression of
mitochondrial genes, resulting in lower rates of ox-phos and decreased NAD+/NADH in the
mitochondria of muscle tissue of old mice 18. Taken together, multiple mechanisms may be
involved in the regulation of the NAD redox states during ageing which may be tissue and context
dependent.
In addition to ageing, altered NAD pools and NAD+/NADH ratios have been observed in
metabolic disease, cellular senescence and in cancer 25–27. In obese conditions when the
NAD+/NADH ratio is low, it inhibits the activity of dehydrogenase enzymes and can lead to
reductive stress, which has been shown to further promote metabolic disease and insulin
resistance 19. In ageing, the physiological consequences of lower NAD+/NADH ratios, and the role
of reductive stress in ageing-related disease is unclear 28. However, interventions that promote
healthy ageing such as caloric restriction and supplementation with NAD+ precursors (NR/NMN)
have been shown to increase the NAD+/NADH ratio 18,29–31. Thus, it is tempting to speculate that
restoring NAD+/NADH ratios to a healthy level may be beneficial to promoting healthy ageing.
However, what this proper ratio is, is still unclear, and it is likely to be cell specific and dependent
on the metabolic demands of each cell type.
One limitation in studying the role of NAD+/NADH redox balance in vitro and in vivo is the
lack of tools to reliably measure and quantify NAD+/NADH ratio in subcellular compartments over
time and in a high throughput manner. However, as discussed in “techniques to study NAD+
metabolism” these barriers have been broken the last few years with new technology such as
NAD(P)-Snifits 7, SoNar 25 and biosensor with a bipartite NAD+- binding domain 2, that allows the
relative quantification of NAD+/NADH and NADP+/NADPH in live cells and in different subcellular
fractions under varying conditions. These tools will be of critical importance to further our
knowledge on the cellular mechanisms that regulate the NAD+/NADH balance and how this
balance in turn influences other cellular functions that impact ageing.
Lastly, in addition to tools to measure the NAD+/NADH ratio, another recent genetic tool
includes the use of the Lactobacillus brevis (Lb)NOX, a bacterial water forming NADH oxidase,
that when expressed in mammalian cells can increase the NAD+/NADH ratio 28, increasing the
flux and directionality of multiple metabolic pathways. Recent work using this tool to tissue
specifically increase the NAD+/NADH ratio in hepatocytes was able to lower reductive stress in
the livers of mice fed a high fat diet and improve glucose metabolism and insulin resistance 27,28.
This data suggests reductive stress not only occurs during metabolically stressful conditions such
as obesity and ageing, but is also causative in promoting metabolic and mitochondrial dysfunction,
and disease. Thus, targeting the NAD+/NADH redox balance may be a potential therapeutic target
to treat metabolic diseases including ageing. Thus, future work using this tool, along with the
others mentioned above, to tissue specifically measure and impact NAD(P) redox states and how
they influence metabolic health and lifespan in ageing mice will be exciting to follow.

Regulation of DNA repair response and genomic stability


A key hallmark of ageing includes a reduced capacity to repair DNA damage and an
accumulation of DNA mutations and genomic instability 32. Excess DNA damage or unresolved
DNA damage, such as that occur during ageing or progeria, results in prolonged PARP1 activation
and depletion of NAD pools 33. Decreased NAD pools from PARP1 activation have downstream
consequences and affect the activity of other NAD+-dependent cellular processes and NAD+
enzymes, including SIRT1, resulting in reduced DNA repair capacity, reduced mitophagy and
altered mitochondrial metabolism and homeostasis 33–35. Therefore, restoring NAD+ levels has
emerged as a potential therapeutic approach to restore or mitigate the negative consequences of
hyper PARP1 activation in response to DNA damage during ageing or other DNA-damaging
insults. In this section, we will explain the emerging evidence for the role of NAD+ in the DNA
damage response and its consequences.
In support of a major role for NAD+ in DNA repair during ageing, repletion with NAD+
precursors, such as NMN and NR, or inhibition of the major NAD+ consumers, such as CD38, are
effective in extending the healthspan and lifespan of many progeroid models, including ataxia-
telangiectasia, Cockayne syndrome and p44+/+ mice 21,36,37. Additionally, NR and NMN treatment
help maintain telomere length in mice lacking telomerase 38 and promote DNA repair in mice with
atherosclerotic lesions, radiation exposure, and even in naturally old wild-type mice, leading to
reduced DNA damage and decreased numbers of senescent cells 39–41. Thus, increasing NAD+
levels in cells experiencing high levels of DNA damage (such as ageing cells) may support DNA
repair and overall genomic stability. However, NADs role in the DNA damage response seems to
be much more complicated than simply acting as a co-substrate for PARP1. For example, when
NAD+ levels are low, PARP1 is bound and sequestered by deleted in breast cancer 1 (DBC1)
protein, which restricts PARP1 activity 39. Interestingly, DBC1 also inhibits SIRT1 activity 42,43, but
this interaction seems to be independent of NAD+ 39. Thus, these results show that, in addition to
NAD’s role as a coenzyme and cosubstrate, NAD+ facilitates protein-protein interactions that
influence DNA repair. In support of this, NAD+ facilitates the interaction of the DNA repair protein
HPF1 and PARP1/2, which completes the PARP activation site and is essential for substrate
PARylation and the DNA repair response 44. These results suggest how decreased NAD+ levels
in ageing leads to diminished PARP1 activity and DNA repair, and how targeting NAD+
consumption routes or administration of NAD+ precursors may promote more efficient DNA repair
mechanisms. While the clinical benefits of targeting NAD+ metabolism remain unclear in humans,
it may be an effective therapeutic route for treating patients with DNA repair defects and progeroid
diseases.
Lastly, sirtuins play indirect and direct roles in genome stability and DNA repair. For
example, the mitochondrial localized SIRT3 suppresses production of mitochondrial derived ROS,
an endogenous source of DNA damage 45–47. Whereas the nuclear localized sirtuins 1, 6, and 7
are integral parts of the DNA repair response to double-stranded breaks. For example, mice
lacking SIRT1, SIRT6, and SIRT7 exhibit genome instability and die within a few weeks after birth
48–50
. Conversely, mice overexpressing these sirtuins have increased healthspan and lifespan 51,52.
SIRT1 is activated by DNA damage and promotes DNA repair via nuclear excision repair 53–55. In
agreement with the role of SIRT1 in regulating genome stability, mice overexpressing SIRT1 are
protected from cancer 56. Similarly, SIRT6 is activated by DNA damage and helps in the repair of
double-stranded breaks by deacetylating histones, recruiting chromatin remodelers, and
promoting PARP1 activation 48,57–59. Interestingly, recent work found that rodent lifespan is
positively correlated with the increased efficiency of SIRT6-dependent repair of double-stranded
breaks 60. Additionally, L1 retrotransposable elements are upregulated during cellular senescence
and are emerging as a key source of DNA damage and genomic instability during ageing 61.
Interestingly, SIRT6 represses L1 retrotransposons by ribosylating the KRAB-associated protein
1 (KAP1), which promotes viral and retrotransposon latency 62. In agreement with this, mice
lacking SIRT6 have shortened lifespan and healthspan that can be extended twofold by treating
SIRT6 KO mice with nucleoside RT inhibitors, antiviral drugs that reduce L1-dependent DNA
damage and genomic instability 63. These studies show that a major function of sirtuins is to
maintain genomic stability and promote DNA repair.
Sirtuin function is tightly linked to NAD+ levels, however in humans it is still unknown
whether augmenting NAD+ levels during ageing can enhance sirtuin dependent DNA repair
mechanisms in addition to those regulated by PARPs. However, it is tempting to speculate that
increasing NAD+ levels will provide therapeutic benefits to human healthspan by activating
multiple downstream NAD+-dependent DNA protective processes discussed above and/or DNA
repair pathways. Thus, an open question in the ageing field is whether NAD+ boosting
supplements can be used as a way to prevent the accumulation of genotoxic stress in the clinic.
Future work untangling how NAD+ influences DNA repair mechanisms will be of utmost
importance to better understand how we can more efficiently target NAD+ metabolism and NAD+-
dependent proteins, such as sirtuins and PARPs, to treat ageing and other diseases associated
with DNA damage.

Epigenetic regulation
The ageing process is characterized by major epigenetic changes, also known as “epigenetic
drift” 64,65 , including reduced global heterochromatin 66,67, changes in histone modification pattern
68
and global DNA hypomethylation 69,70,71, which together affect the genomic stability, chromatin
state and gene expression.
The discovery that enzymes involved in epigenetic regulation, such as the histone
deacetylases sirtuins, use NAD+ as a cofactor highlighted the pivotal role of this metabolite in the
crosstalk between metabolic state of the cell and epigenetic regulation of gene expression 72,73.
Increased NAD+ levels in response to caloric restriction (CR) 74,75 and physical exercise 76,
increase sirtuin activity which transduce the signal via deacetylation of specific histone residues
(e.g., H3K9, H3K14 and H4K16) 77 and transcription factors (e.g., p53, NF-kB, PGC1a and FOXO
) 78–80 resulting in beneficial anti-ageing effects. An extensive body of evidence shows in particular
the role of SIRT1 in modulating the age-related epigenetic landscape and the potential of targeting
this enzyme with activating compounds, such as resveratrol, which reduce genomic instability and
histone acetylation (e.g., H3K9 H4K16, and H3K56) and increased heterochromatin formation
55,81,82
. Moreover, the circadian regulation of NAD+ levels orchestrated by the clock machinery
CLOCK:BMAL1 and SIRT1 reveal a circadian regulation of the SIRT1 epigenetic targets such as
H3K9 and H3K14 at multiple loci 83 83,84. Beside the established role of SIRT1 as a NAD+-
dependent biosensor for the nutrient state of the cell and epigenetic regulator, a growing body of
evidence shows the involvement of other members of the sirtuin family including SIRT2 85, -6 86,87
and -7 88 (more extensively reviewed here 89). Moreover, accumulating evidence suggests the
involvement of PARPs and PARylation in epigenetic regulation, in particular in heterochromatin
integrity and architecture (reviewed in 90), however its role in ageing is still unclear. Decreased
PARylation has been associated with hypoacetylation 91 possibly due to the reduced NAD+
availability for SIRT1 mediated by PARP1 92.
One key concept related to age-related epigenetic changes is the relocalization of
chromatin modifiers (RCM) hypothesis, suggesting that in ageing cells the epigenetic modifiers
recruited in the repair process do not return to their original site once the repair process is
completed 93. This results in an altered landscape of these chromatin modifiers which affects the
gene expression profile and contributes to loss of cell identity and senescence 94. The involvement
of sirtuins in the RCM process has been demonstrated in yeast 55,58, however, the contribution of
the RCM to mammalian ageing and the role of NAD+ metabolism is still largely unexplored.
One exciting area in the ageing/epigenetics field is the discovery that changes in DNA
methylation occurs at a predictable rate, which can be used as a biological clock (known as the
Horvath Clock) to predict and determine the rate of ageing in humans 95. Interestingly, a recent
small clinical study, consisting of 9 middle aged men were given a cocktail of three drugs
(recombinant human growth hormone, dehydroepiandrosterone (DHEA) and metformin) for one
year to try to reverse immunosenescence and ageing as predicted by the Horvath clock 96.
Interestingly, analysis of peripheral blood immune cells in these patients showed that this cocktail
was able to reduce the DNA methylation clock on average by 2.5 years, and also able to reduce
the expression of CD38 in peripheral blood monocytes in these patients suggesting this reversal
may be dependent on NAD+ levels. Thus, this exciting study hints, but does not prove, that
targeting NAD+ metabolism may be used to reverse specific age-related epigenetic features.
Therefore, future studies are required to advance our understanding of its real therapeutic
potential.
Taken together, accumulating evidence indicates the beneficial effect of increased NAD+
levels via upregulation of epigenetic regulators (e.g., SIRT1) in reversing age-related epigenetic
changes and increasing healthspan 97–100 . Thus, it remains to be determined whether other NAD+
boosting strategies such as supplementation with NAD+ precursors could have the same
beneficial effects.

Autophagy
Autophagy is a highly conserved catabolic process that mediates the degradation of defective
organelles and protein aggregates, particularly during conditions when nutrients are low, for
recycling in the lysosomes, and is thought to promote beneficial effects for cellular homeostasis
and more largely to organismal fitness. However, defective autophagy is a common signature of
ageing and contributes to age-related diseases, including metabolic and cardiac disorders and
neurodegeneration 101,102,103. Autophagy is controlled by multiple nutritional and stress-related
cues and by a set of highly conserved autophagy (Atg) proteins 104. The reasons why autophagy
decreases with age are unclear but down-regulation of these Atg genes such as Atg5 and Atg7
have been detected in human tissues during normal ageing 105.
Increasing data indicate that NAD+/SIRT1 signaling pathway increases autophagic flux by
regulating the molecular mechanisms of autophagy induction and initiation 106 through the
deacetylation of several ATG proteins, including ATG7, ATG6 107 and LC3 108 . In addition, the
NAD+/SIRT1 pathway also induces autophagy through the deacetylation of FOXO1 and
consequently the upregulation of Rab7 in cardiomyocytes 109 and the activation of AMPK in
primary neurons, leading to inhibition of mTOR 110 . Thus, these data link multiple nutrient sensing
pathways, including those of NAD+ via SIRT1, to control of autophagy. In support of NAD levels
influencing autophagy, it has been demonstrated that cardiac CD38 expression is elevated under
hypoxia/Ischemia injury, leading to reduced NAD+ levels, and cardiac dysfunction by inhibiting
autophagic flux via NAD+-dependent Rab7 and non-NAD+-dependent PLEKHM1 downregulation
111
. Thus, autophagy is a major cellular mechanism that is directly regulated by NAD+ levels and
can be potentially improved or made more efficient in ageing patients by targeting NAD+
metabolism pathways.
Mitophagy is a selective autophagy mechanism that plays a major role in specific
recognition and degradation of damaged or ROS-producing mitochondria. NAD+ depletion has
been associated to loss of mitophagy and mitochondrial dysfunction in several studies of
premature ageing syndromes including Werner syndrome 112 Xeroderma Pigmentosum 33,
Cockayne syndrome 21 and Ataxia-telangiectasia 113 due to PARP1 hyperactivation induced by
DNA damage and the impairment of sirtuin activity. Moreover, it was demonstrated that
supplementation with NAD+ precursors such as NR and NMN promotes the sirtuin dependent
mitochondrial recycling and clearance of protein aggregates in relevant models of
neurodegenerative disorders including AD 114,115 and PD 116 via increased TFEB- and FOXO-
dependent expression of autophagy/mitophagy genes. Lastly, the NAD+-consuming enzyme
SARM1 promotes mitophagy by forming a complex with PTEN-induced putative kinase 1 (PINK1)
and TRAF6 (TNF receptor associated factor 6) that stabilizes PINK1 on depolarized mitochondria
supporting a link between the deregulation of complex formation and PD pathogenesis 117. Thus,
multiple NAD+ dependent pathways and enzymes converge to control mitophagy, suggesting
there are multiple therapeutic approaches targeting NAD+ metabolism that may promote
mitophagy and better ageing.
Supplementary Box 3. NAD+ in cancer

Ageing is one of the main risk factors for several types of cancer, which are characterized by age-
related pathophysiological processes regulated by NAD+ levels, including oxidative stress and
DNA damage 118 , metabolic dysfunction 119, cellular senescence 120 and inflammation 121,122. NAD+
metabolism is emerging as a potential target for cancer treatment since many NAD+-related
enzymes are dysregulated in carcinogenesis.
Both iNAMPT and eNAMPT are overexpressed in numerous malignancies 123 including
prostate 124, breast 125 and colorectal cancers 126 resulting in increased NAD+ levels which promote
cell proliferation and cancer initiation and progression. Several NAMPT inhibitors (i.e. CHS828 127
and FK866 128) have been developed and showed promising antineoplastic activity in vitro but
very low efficacy and high toxicity in clinical trials 129. The role of other NAD+ biosynthetic enzymes
is less clear but NMNAT2 overexpression in cancer has been reported 130,131. Overall, in contrast
with the pathological role of NAD+ decline in ageing, this evidence suggests the detrimental role
of an upregulated NAD+ biosynthesis in cancer.
The role of NAD+-consuming enzymes in cancer may vary from tumour promoters to
tumour suppressors. Evidence showing that sirtuins may both promote and suppress
tumorigenesis targeting multiple pathways has been extensively reported and recently reviewed
132
. For example, while SIRT1 overexpression is associated with poor disease progression and
poor prognosis in colorectal cancer 133, its expression in non-small cell lung cancer seems to
attenuate tumour progression via NF-kB signaling regulation 134. SIRT1 was also reported to
downregulate several tumour suppressors including p53 135,136 and PTEN 137, and stabilize
oncogenes such as MYCN 138.
Since genomic instability and DNA damage are key underlying factors in most types of
cancers 139 but also side effects of chemotherapeutic agents, the role of PARPs in DNA repair
and cancer has been extensively studied. Remarkably, targeting PARPs shows encouraging
results in breast and ovarian cancers with BRCA mutations characterized by a defective
homologous recombination repair system and several PARP inhibitors were developed and FDA-
approved 140–142. In BRCA deficient cancers, the inhibition of PARP activity promotes synthetic
lethality 143,144 via different mechanisms including “PARP trapping” 145,146, however the role of
NAD+ level in this process is unclear.
Lastly, CD38 is overexpressed in several hematologic malignancies, and its expression is
often associated with poor prognosis 147–150,151. One of the blood cancers with the highest and
most frequent expression of CD38 is in patients with multiple myeloma (MM), where CD38 is
highly expressed on malignant plasma cells 147–150. As a result, in the last decade, several anti-
CD38 antibodies were developed to efficiently and selectively target CD38 in MM patients. So far,
the clinical results have been extremely promising 152–154. Interestingly, CD38 KO in the ARH1
multi-tumour mouse model shows higher NAD+ levels and strong inhibition of tumour development
151
. Recent evidence shows the detrimental role of CD38 in suppressing anti-tumour effector
functions mediated by several immune populations. Th1/17 hybrid T cells with high levels of
NAD+-dependent SIRT1 and reduced CD38 expression exhibited enhanced anti-tumour activity,
therefore suggesting that blocking CD38 augments adoptive T-cell transfer in cancer patients 155.
Moreover, a new role of CD38 was proposed in the CD8-mediated immune control of tumours
treated with the checkpoint inhibitors PD-1/PD-L1 antagonist antibodies 156,157. After PD-1/PD-L1
blockade therapy, CD38 is overexpressed by CD8 T cells, leading to CD8 T-cell exhaustion, poor
antitumour immunity and ultimately resistance. Interestingly, targeting CD38 on this dysfunctional
immune population enhanced tumour control in in vivo models with poor therapeutic responses
to PD-1/PD-L1 blockade therapy 156, 157. The resistance mechanisms driven by CD38 are still
unclear; however, they seem to involve a perturbation of the adenosine receptor signaling 156,157.
Overall, CD38 may represent an innovative target in combination with checkpoint inhibitors to
improve the T-cell-mediated immune response to tumours. Unlike the case of CD38, data on the
role of CD157 in diseases are still limited. In hematologic malignancies, CD157 has a restricted
pattern of expression, and it’s mostly associated with ovarian cancer158 and acute myeloid
leukemia (AML) where an anti-CD157-based therapy shows encouraging results159.

Supplementary Box 4. Techniques to study NAD+ metabolism

Implementing current analytical methods to detect NAD+-related metabolites is one of the main
topics in the NAD+ field. Over the past decades, different techniques have been employed
including NAD+/NADH enzymatic cycling assays 160, a multitude of high-performance liquid
chromatography (HPLC) based-methods, such as HPLC with UV 160,161 or fluorometric detection
162
, and more sophisticated LC/MS/MS 163 or HPLC/MALDI/MS 164 systems. Despite the
advantages achieved in terms of sensitivity and accuracy with the mass-spectrometry based
technologies, quantification of NAD(P)+ (as well as detection of its reduced form) in different
cellular compartments remains a challenge due to the complexity of subcellular fractionations,
instability of the redox state during metabolite extraction, and suboptimal metabolite isolation
procedures.
However, the recent development of semisynthetic fluorescent biosensors, such as BRET-
based biosensors 6, NAD(P)-Snifits 7, SoNar 25 and biosensor with a bipartite NAD+- binding
domain2, allows the relative quantification of free NAD+/NADH and NADP+/NADPH in live cells
and in different subcellular fractions. Nevertheless, this method has been established for in vitro
experiments and further studies are needed to translate it in animal models.
Estimating NAD+ synthesis and consumption rates based on unlabeled analyte
concentrations or biochemical assays is inadequate to narrow down the different pathways of
generating and salvaging NAD+. Although tracing experiments with decaying isotopically labeled
NAD+ have been reported since the 80s 165 , the advance in mass spectrometry nowadays allows
flux measurements using stable isotopes, with quantitative measurement of unlabeled and
labeled forms of different NAD+-related metabolites.
Single labeled isotopic forms (2H and 13C) of nicotinamide, nicotinic acid, NAD+ and
tryptophan are commercially available whereas a new era of dedicated enzymatic or chemical
syntheses of doubly labelled NAD+-related metabolites (2H, 18O, 13C and 15N isotopes) in which
one heavier isotope is incorporated on the nicotinamide ring and one heavier isotope on the ribose
moieties is taking place. Indeed, this type of doubly labeling provides valuable information
regarding flows of metabolites in the NAD+ biosynthetic and turn-over pathways. For instance,
doubly labeled NMN or NR were employed in vitro and in vivo experiments to elucidate the direct
uptake and incorporation into NAD+ 166,167,168 as well as isotopic labeling experiments using
nicotinic acid riboside were performed to discern the ability of mitochondria to directly import intact
NAD+ 5.
Supplementary Table 1. NAD+ and hallmarks of ageing

Hallmarks of ageing Pathways and conditions References


affected by NAD+ impairment
or NAD+-dependent enzymes
33,169,170,39
Genomic instability In case of premature ageing
diseases (XPA, CS, A-T), mutation in
DNA repair mechanisms leads to
PARP1 hyperactivation and
reduction in sirtuin activity.
Age-dependent NAD+ decline
decreases PARP1 activity by
promoting the interaction with
DBC1.

38
Telomere attrition
Telomere shortening in livers of
TERT KO mice leads to a p53 -
dependent repression of all seven
sirtuins.

171
Epigenetic alterations SIRT1-dependent gene regulation
through histone modification, DNA
methylation, and the modulation of
chromatin structure is affected by
NAD+ decline during ageing

172, 113,169,173
Loss of proteostasis In AD models, mitophagy is
enhanced by NAD+ boosting
strategies via a mechanism
mediated by pink1, pdr-1, and dct-1.

In mice and worm, Sirtuin-mediated


UPRmt response and FOXO
signaling is activated by NAD+
augmentation or PARP inhibition

Deregulated nutrient sensing NAD+ decrease affects 174, 175, 176

AMPK/sirtuin/PGC1α pathways.

24, 18
Mitochondrial dysfunction
Age-dependent CD38 increase
affects the availability of NAD+ to
mitochondrial enzymes including
SIRT3, leading to impairment of
oxygen consumption
In worm and mice, decline in NAD+
and accumulation of HIF-1α causes
loss of OXPHOS subunits via
impaired SIRT1-PGC-1a signaling

177,178
Cellular senescence Senescence associated secretory
proteins (SASP) secreted by
senescent cells induce CD38-
NADase activity in non-senescent
cells (macrophages or endothelial
cells) that ultimately lead to tissue
NAD+ decline

Stem cell exhaustion NAD+ supplementation rejuvenates 41,179–181

aged gut, muscle and hematopoietic


stem cell pools by enhancing sirtuin
1 activity or mitochondria clearance

Altered intercellular The effect of NAD+ augmentation on 114, 115, 182

communication inflammation has also been


demonstrated in age-associated
disease including diabetes and AD
by reducing pro-inflammatory
cytokines and major inflammasome
NLRP3
References:

1. Wanders, R. J. A., Waterham, H. R. & Ferdinandusse, S. Metabolic Interplay between

Peroxisomes and Other Subcellular Organelles Including Mitochondria and the Endoplasmic

Reticulum. Front Cell Dev Biol 3, 83 (2015).

2. Cambronne, X. A. et al. Biosensor reveals multiple sources for mitochondrial NAD+. Science

352, 1474–1477 (2016).

3. Luongo, T. S. et al. SLC25A51 is a mammalian mitochondrial NAD transporter. Nature (2020)

doi:10.1038/s41586-020-2741-7.

4. Kory, N. et al. MCART1/SLC25A51 is required for mitochondrial NAD transport. Sci Adv 6,

(2020).

5. Davila, A. et al. Nicotinamide adenine dinucleotide is transported into mammalian

mitochondria. Elife 7, (2018).

6. Yu, Q. et al. A biosensor for measuring NAD levels at the point of care. Nature Metabolism

vol. 1 1219–1225 (2019).

7. Sallin, O. et al. Semisynthetic biosensors for mapping cellular concentrations of nicotinamide

adenine dinucleotides. Elife 7, (2018).

8. Adriouch, S. et al. NAD Released during Inflammation Participates in T Cell Homeostasis by

Inducing ART2-Mediated Death of Naive T Cells In Vivo. The Journal of Immunology vol. 179

186–194 (2007).

9. Haag, F. et al. Extracellular NAD and ATP: Partners in immune cell modulation. Purinergic

Signal. 3, 71–81 (2007).

10. Breen, L. T., Smyth, L. M., Yamboliev, I. A. & Mutafova-Yambolieva, V. N. beta-NAD is a

novel nucleotide released on stimulation of nerve terminals in human urinary bladder detrusor

muscle. Am. J. Physiol. Renal Physiol. 290, F486–95 (2006).

11. Mutafova-Yambolieva, V. N. et al. Beta-nicotinamide adenine dinucleotide is an inhibitory

neurotransmitter in visceral smooth muscle. Proc. Natl. Acad. Sci. U. S. A. 104, 16359–16364
(2007).

12. Mottahedeh, J. et al. CD38 is methylated in prostate cancer and regulates extracellular NAD.

Cancer Metab 6, 13 (2018).

13. O’Reilly, T. & Niven, D. F. Levels of nicotinamide adenine dinucleotide in extracellular body

fluids of pigs may be growth-limiting for Actinobacillus pleuropneumoniae and Haemophilus

parasuis. Can. J. Vet. Res. 67, 229–231 (2003).

14. Brunnbauer, P. et al. The nanomolar sensing of nicotinamide adenine dinucleotide in human

plasma using a cycling assay in albumin modified simulated body fluids. Sci. Rep. 8, 16110

(2018).

15. Bruzzone, S., Guida, L., Zocchi, E., Franco, L. & De Flora A. Connexin 43 hemi channels

mediate Ca2+-regulated transmembrane NAD+ fluxes in intact cells. FASEB J. 15, 10–12

(2001).

16. Moreschi, I. et al. Extracellular NAD Is an Agonist of the Human P2Y11Purinergic Receptor

in Human Granulocytes. Journal of Biological Chemistry vol. 281 31419–31429 (2006).

17. Adriouch, S., Haag, F., Boyer, O., Seman, M. & Koch-Nolte, F. Extracellular NAD : a danger

signal hindering regulatory T cells. Microbes and Infection vol. 14 1284–1292 (2012).

18. Gomes, A. P. et al. Declining NAD(+) induces a pseudohypoxic state disrupting nuclear-

mitochondrial communication during aging. Cell 155, 1624–1638 (2013).

19. Heer, C. D. & Brenner, C. Letting off electrons to cope with metabolic stress. Nature

Metabolism (2020) doi:10.1038/s42255-020-0207-8.

20. Zhu, X.-H., Lu, M., Lee, B.-Y., Ugurbil, K. & Chen, W. In vivo NAD assay reveals the

intracellular NAD contents and redox state in healthy human brain and their age

dependences. Proceedings of the National Academy of Sciences vol. 112 2876–2881 (2015).

21. Scheibye-Knudsen, M. et al. A high-fat diet and NAD(+) activate Sirt1 to rescue premature

aging in cockayne syndrome. Cell Metab. 20, 840–855 (2014).

22. Desdín-Micó, G. et al. T cells with dysfunctional mitochondria induce multimorbidity and
premature senescence. Science (2020) doi:10.1126/science.aax0860.

23. Braidy, N. et al. Age related changes in NAD+ metabolism oxidative stress and Sirt1 activity

in wistar rats. PLoS One 6, e19194 (2011).

24. Camacho-Pereira, J. et al. CD38 Dictates Age-Related NAD Decline and Mitochondrial

Dysfunction through an SIRT3-Dependent Mechanism. Cell Metab. 23, 1127–1139 (2016).

25. Zhao, Y. et al. SoNar, a Highly Responsive NAD /NADH Sensor, Allows High-Throughput

Metabolic Screening of Anti-tumor Agents. Cell Metabolism vol. 21 777–789 (2015).

26. Nacarelli, T. et al. NAD metabolism governs the proinflammatory senescence-associated

secretome. Nat. Cell Biol. 21, 397–407 (2019).

27. Goodman, R. P. et al. Hepatic NADH reductive stress underlies common variation in

metabolic traits. Nature (2020) doi:10.1038/s41586-020-2337-2.

28. Titov, D. V. et al. Complementation of mitochondrial electron transport chain by manipulation

of the NAD+/NADH ratio. Science 352, 231–235 (2016).

29. Cantó, C. et al. The NAD(+) precursor nicotinamide riboside enhances oxidative metabolism

and protects against high-fat diet-induced obesity. Cell Metab. 15, 838–847 (2012).

30. Lin, S.-J., Ford, E., Haigis, M., Liszt, G. & Guarente, L. Calorie restriction extends yeast life

span by lowering the level of NADH. Genes Dev. 18, 12–16 (2004).

31. Sims, C. A. et al. Nicotinamide mononucleotide preserves mitochondrial function and

increases survival in hemorrhagic shock. JCI Insight 3, (2018).

32. López-Otín, C., Blasco, M. A., Partridge, L., Serrano, M. & Kroemer, G. The Hallmarks of

Aging. Cell vol. 153 1194–1217 (2013).

33. Fang, E. F. et al. Defective mitophagy in XPA via PARP-1 hyperactivation and NAD(+)/SIRT1

reduction. Cell 157, 882–896 (2014).

34. Croteau, D. L., Fang, E. F., Nilsen, H. & Bohr, V. A. NAD in DNA repair and mitochondrial

maintenance. Cell Cycle 16, 491–492 (2017).

35. Fang, E. F. et al. Nuclear DNA damage signalling to mitochondria in ageing. Nat. Rev. Mol.
Cell Biol. 17, 308–321 (2016).

36. Fang, E. F. et al. NAD Replenishment Improves Lifespan and Healthspan in Ataxia

Telangiectasia Models via Mitophagy and DNA Repair. Cell Metab. 24, 566–581 (2016).

37. Tarragó, M. G. et al. A Potent and Specific CD38 Inhibitor Ameliorates Age-Related Metabolic

Dysfunction by Reversing Tissue NAD Decline. Cell Metab. 27, 1081–1095.e10 (2018).

38. Amano, H. et al. Telomere Dysfunction Induces Sirtuin Repression that Drives Telomere-

Dependent Disease. Cell Metab. 29, 1274–1290.e9 (2019).

39. Li, J. et al. A conserved NAD binding pocket that regulates protein-protein interactions during

aging. Science 355, 1312–1317 (2017).

40. Haemmig, S. et al. Long noncoding RNA SNHG12 integrates a DNA-PK–mediated DNA

damage response and vascular senescence. Science Translational Medicine vol. 12

eaaw1868 (2020).

41. Zhang, H. et al. NAD+ repletion improves mitochondrial and stem cell function and enhances

life span in mice. Science 352, 1436–1443 (2016).

42. Zhao, W. et al. Negative regulation of the deacetylase SIRT1 by DBC1. Nature vol. 451 587–

590 (2008).

43. Kim, J.-E., Chen, J. & Lou, Z. DBC1 is a negative regulator of SIRT1. Nature vol. 451 583–

586 (2008).

44. Suskiewicz, M. J. et al. HPF1 completes the PARP active site for DNA damage-induced ADP-

ribosylation. Nature (2020) doi:10.1038/s41586-020-2013-6.

45. Qiu, X., Brown, K., Hirschey, M. D., Verdin, E. & Chen, D. Calorie restriction reduces oxidative

stress by SIRT3-mediated SOD2 activation. Cell Metab. 12, 662–667 (2010).

46. Tao, R. et al. Sirt3-mediated deacetylation of evolutionarily conserved lysine 122 regulates

MnSOD activity in response to stress. Mol. Cell 40, 893–904 (2010).

47. Haigis, M. C., Deng, C.-X., Finley, L. W. S., Kim, H.-S. & Gius, D. SIRT3 is a mitochondrial

tumor suppressor: a scientific tale that connects aberrant cellular ROS, the Warburg effect,
and carcinogenesis. Cancer Res. 72, 2468–2472 (2012).

48. Mostoslavsky, R. et al. Genomic instability and aging-like phenotype in the absence of

mammalian SIRT6. Cell 124, 315–329 (2006).

49. McBurney, M. W. et al. The Mammalian SIR2 Protein Has a Role in Embryogenesis and

Gametogenesis. Molecular and Cellular Biology vol. 23 38–54 (2003).

50. Vazquez, B. N. et al. SIRT7 promotes genome integrity and modulates non-homologous end

joining DNA repair. EMBO J. 35, 1488–1503 (2016).

51. Kanfi, Y. et al. The sirtuin SIRT6 regulates lifespan in male mice. Nature 483, 218–221

(2012).

52. Satoh, A. et al. Sirt1 extends life span and delays aging in mice through the regulation of Nk2

homeobox 1 in the DMH and LH. Cell Metab. 18, 416–430 (2013).

53. Ming, M. et al. Regulation of global genome nucleotide excision repair by SIRT1 through

xeroderma pigmentosum C. Proc. Natl. Acad. Sci. U. S. A. 107, 22623–22628 (2010).

54. Fan, W. & Luo, J. SIRT1 regulates UV-induced DNA repair through deacetylating XPA. Mol.

Cell 39, 247–258 (2010).

55. Oberdoerffer, P. et al. SIRT1 redistribution on chromatin promotes genomic stability but alters

gene expression during aging. Cell 135, 907–918 (2008).

56. Herranz, D. et al. Sirt1 improves healthy ageing and protects from metabolic syndrome-

associated cancer. Nat. Commun. 1, 3 (2010).

57. McCord, R. A. et al. SIRT6 stabilizes DNA-dependent protein kinase at chromatin for DNA

double-strand break repair. Aging 1, 109–121 (2009).

58. Mao, Z. et al. SIRT6 promotes DNA repair under stress by activating PARP1. Science 332,

1443–1446 (2011).

59. Toiber, D. et al. SIRT6 Recruits SNF2H to DNA Break Sites, Preventing Genomic Instability

through Chromatin Remodeling. Molecular Cell vol. 51 454–468 (2013).

60. Tian, X. et al. SIRT6 Is Responsible for More Efficient DNA Double-Strand Break Repair in
Long-Lived Species. Cell 177, 622–638.e22 (2019).

61. De Cecco, M. et al. Author Correction: L1 drives IFN in senescent cells and promotes age-

associated inflammation. Nature 572, E5 (2019).

62. Van Meter, M. et al. SIRT6 represses LINE1 retrotransposons by ribosylating KAP1 but this

repression fails with stress and age. Nat. Commun. 5, 5011 (2014).

63. Simon, M. et al. LINE1 Derepression in Aged Wild-Type and SIRT6-Deficient Mice Drives

Inflammation. Cell Metab. 29, 871–885.e5 (2019).

64. Mendelsohn, A. R. & Larrick, J. W. Epigenetic Drift Is a Determinant of Mammalian Lifespan.

Rejuvenation Research vol. 20 430–436 (2017).

65. Lara*, E., Calvanese*, V. & Fraga, M. F. Epigenetic Drift and Aging. Epigenetics of Aging

257–273 (2010) doi:10.1007/978-1-4419-0639-7_14.

66. Scaffidi, P. Lamin A-Dependent Nuclear Defects in Human Aging. Science vol. 312 1059–

1063 (2006).

67. Zhang, W. et al. Aging stem cells. A Werner syndrome stem cell model unveils

heterochromatin alterations as a driver of human aging. Science 348, 1160–1163 (2015).

68. Ryu, S. H., Kang, K., Yoo, T., Joe, C. O. & Chung, J. H. Transcriptional repression of repeat-

derived transcripts correlates with histone hypoacetylation at repetitive DNA elements in

aged mice brain. Exp. Gerontol. 46, 811–818 (2011).

69. Berdyshev, G. D., Korotaev, G. K., Boiarskikh, G. V. & Vaniushin, B. F. [Nucleotide

composition of DNA and RNA from somatic tissues of humpback and its changes during

spawning]. Biokhimiia 32, 988–993 (1967).

70. Christensen, B. C. et al. Aging and environmental exposures alter tissue-specific DNA

methylation dependent upon CpG island context. PLoS Genet. 5, e1000602 (2009).

71. Horvath, S. & Raj, K. DNA methylation-based biomarkers and the epigenetic clock theory of

ageing. Nat. Rev. Genet. 19, 371–384 (2018).

72. Gut, P. & Verdin, E. The nexus of chromatin regulation and intermediary metabolism. Nature
vol. 502 489–498 (2013).

73. Imai, S., Armstrong, C. M., Kaeberlein, M. & Guarente, L. Transcriptional silencing and

longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature 403, 795–800

(2000).

74. Lin, S. J., Defossez, P. A. & Guarente, L. Requirement of NAD and SIR2 for life-span

extension by calorie restriction in Saccharomyces cerevisiae. Science 289, 2126–2128

(2000).

75. Moroz, N. et al. Dietary restriction involves NAD -dependent mechanisms and a shift toward

oxidative metabolism. Aging Cell vol. 13 1075–1085 (2014).

76. Koltai, E. et al. Exercise alters SIRT1, SIRT6, NAD and NAMPT levels in skeletal muscle of

aged rats. Mech. Ageing Dev. 131, 21–28 (2010).

77. Bosch-Presegué, L. & Vaquero, A. Sirtuin-dependent epigenetic regulation in the

maintenance of genome integrity. FEBS J. 282, 1745–1767 (2015).

78. Brunet, A. et al. Stress-dependent regulation of FOXO transcription factors by the SIRT1

deacetylase. Science 303, 2011–2015 (2004).

79. Rodgers, J. T. et al. Nutrient control of glucose homeostasis through a complex of PGC-1α

and SIRT1. Nature vol. 434 113–118 (2005).

80. Vaziri, H. et al. hSIR2(SIRT1) functions as an NAD-dependent p53 deacetylase. Cell 107,

149–159 (2001).

81. Chen, Y. et al. Quantitative acetylome analysis reveals the roles of SIRT1 in regulating

diverse substrates and cellular pathways. Mol. Cell. Proteomics 11, 1048–1062 (2012).

82. Vaquero, A. et al. SIRT1 regulates the histone methyl-transferase SUV39H1 during

heterochromatin formation. Nature 450, 440–444 (2007).

83. Nakahata, Y., Sahar, S., Astarita, G., Kaluzova, M. & Sassone-Corsi, P. Circadian control of

the NAD+ salvage pathway by CLOCK-SIRT1. Science 324, 654–657 (2009).

84. Belden, W. J. & Dunlap, J. C. SIRT1 Is a Circadian Deacetylase for Core Clock Components.
Cell vol. 134 212–214 (2008).

85. Serrano, L. et al. The tumor suppressor SirT2 regulates cell cycle progression and genome

stability by modulating the mitotic deposition of H4K20 methylation. Genes & Development

vol. 27 639–653 (2013).

86. Michishita, E. et al. SIRT6 is a histone H3 lysine 9 deacetylase that modulates telomeric

chromatin. Nature vol. 452 492–496 (2008).

87. Michishita, E. et al. Cell cycle-dependent deacetylation of telomeric histone H3 lysine K56 by

human SIRT6. Cell Cycle vol. 8 2664–2666 (2009).

88. Barber, M. F. et al. SIRT7 links H3K18 deacetylation to maintenance of oncogenic

transformation. Nature vol. 487 114–118 (2012).

89. Jing, H. & Lin, H. Sirtuins in epigenetic regulation. Chem. Rev. 115, 2350–2375 (2015).

90. Ciccarone, F., Zampieri, M. & Caiafa, P. PARP1 orchestrates epigenetic events setting up

chromatin domains. Semin. Cell Dev. Biol. 63, 123–134 (2017).

91. Verdone, L. et al. Poly(ADP-Ribosyl)ation Affects Histone Acetylation and Transcription.

PLoS One 10, e0144287 (2015).

92. Cantó, C., Sauve, A. A. & Bai, P. Crosstalk between poly(ADP-ribose) polymerase and sirtuin

enzymes. Mol. Aspects Med. 34, 1168–1201 (2013).

93. Oberdoerffer, P. & Sinclair, D. A. The role of nuclear architecture in genomic instability and

ageing. Nat. Rev. Mol. Cell Biol. 8, 692–702 (2007).

94. Kane, A. E. & Sinclair, D. A. Epigenetic changes during aging and their reprogramming

potential. Crit. Rev. Biochem. Mol. Biol. 54, 61–83 (2019).

95. Horvath, S. DNA methylation age of human tissues and cell types. Genome Biology vol. 14

R115 (2013).

96. Fahy, G. M. et al. Reversal of epigenetic aging and immunosenescent trends in humans.

Aging Cell 18, e13028 (2019).

97. Lamming, D. W. et al. HST2 mediates SIR2-independent life-span extension by calorie


restriction. Science 309, 1861–1864 (2005).

98. Tissenbaum, H. A. & Guarente, L. Increased dosage of a sir-2 gene extends lifespan in

Caenorhabditis elegans. Nature 410, 227–230 (2001).

99. Wood, J. G. et al. Sirtuin activators mimic caloric restriction and delay ageing in metazoans.

Nature 430, 686–689 (2004).

100. Ho, C., van der Veer, E., Akawi, O. & Pickering, J. G. SIRT1 markedly extends replicative

lifespan if the NAD+ salvage pathway is enhanced. FEBS Lett. 583, 3081–3085 (2009).

101. Rubinsztein, D. C., Mariño, G. & Kroemer, G. Autophagy and Aging. Cell vol. 146 682–695

(2011).

102. Barbosa, M. C., Grosso, R. A. & Fader, C. M. Hallmarks of Aging: An Autophagic Perspective.

Front. Endocrinol. 9, 790 (2018).

103. Shirakabe, A., Ikeda, Y., Sciarretta, S., Zablocki, D. K. & Sadoshima, J. Aging and Autophagy

in the Heart. Circ. Res. 118, 1563–1576 (2016).

104. Ohsumi, Y. Historical landmarks of autophagy research. Cell Res. 24, 9–23 (2014).

105. Lipinski, M. M. et al. Genome-wide analysis reveals mechanisms modulating autophagy in

normal brain aging and in Alzheimer’s disease. Proceedings of the National Academy of

Sciences vol. 107 14164–14169 (2010).

106. Lee, I. H. et al. A role for the NAD-dependent deacetylase Sirt1 in the regulation of autophagy.

Proceedings of the National Academy of Sciences vol. 105 3374–3379 (2008).

107. Kerr, J. S. et al. Mitophagy and Alzheimer’s Disease: Cellular and Molecular Mechanisms.

Trends in Neurosciences vol. 40 151–166 (2017).

108. Huang, R. et al. Deacetylation of Nuclear LC3 Drives Autophagy Initiation under Starvation.

Molecular Cell vol. 57 456–466 (2015).

109. Hariharan, N. et al. Deacetylation of FoxO by Sirt1 Plays an Essential Role in Mediating

Starvation-Induced Autophagy in Cardiac Myocytes. Circulation Research vol. 107 1470–

1482 (2010).
110. Wang, P. et al. Induction of autophagy contributes to the neuroprotection of nicotinamide

phosphoribosyltransferase in cerebral ischemia. Autophagy vol. 8 77–87 (2012).

111. Zhang, X. et al. CD38 Causes Autophagic Flux Inhibition and Cardiac Dysfunction Through

a Transcriptional Inhibition Pathway Under Hypoxia/Ischemia Conditions. Front Cell Dev Biol

8, 191 (2020).

112. Fang, E. F. et al. NAD augmentation restores mitophagy and limits accelerated aging in

Werner syndrome. Nat. Commun. 10, 5284 (2019).

113. Fang, E. F. et al. NAD Replenishment Improves Lifespan and Healthspan in Ataxia

Telangiectasia Models via Mitophagy and DNA Repair. Cell Metab. 24, 566–581 (2016).

114. Fang, E. F. et al. Mitophagy inhibits amyloid-β and tau pathology and reverses cognitive

deficits in models of Alzheimer’s disease. Nat. Neurosci. 22, 401–412 (2019).

115. Hou, Y. et al. NAD supplementation normalizes key Alzheimer’s features and DNA damage

responses in a new AD mouse model with introduced DNA repair deficiency. Proc. Natl.

Acad. Sci. U. S. A. 115, E1876–E1885 (2018).

116. Schöndorf, D. C. et al. The NAD+ Precursor Nicotinamide Riboside Rescues Mitochondrial

Defects and Neuronal Loss in iPSC and Fly Models of Parkinson’s Disease. Cell Rep. 23,

2976–2988 (2018).

117. Murata, H., Sakaguchi, M., Kataoka, K. & Huh, N.-H. SARM1 and TRAF6 bind to and stabilize

PINK1 on depolarized mitochondria. Mol. Biol. Cell 24, 2772–2784 (2013).

118. Hoeijmakers, J. H. J. DNA damage, aging, and cancer. N. Engl. J. Med. 361, 1475–1485

(2009).

119. Extermann, M. Metabolic syndrome, aging, and cancer. Crit. Rev. Oncog. 18, 515–529

(2013).

120. Campisi, J. Aging, cellular senescence, and cancer. Annu. Rev. Physiol. 75, 685–705 (2013).

121. Bottazzi, B., Riboli, E. & Mantovani, A. Aging, inflammation and cancer. Semin. Immunol. 40,

74–82 (2018).
122. Lasry, A. & Ben-Neriah, Y. Senescence-associated inflammatory responses: aging and

cancer perspectives. Trends Immunol. 36, 217–228 (2015).

123. Garten, A., Petzold, S., Körner, A., Imai, S.-I. & Kiess, W. Nampt: linking NAD biology,

metabolism and cancer. Trends Endocrinol. Metab. 20, 130–138 (2009).

124. Wang, B. et al. NAMPT overexpression in prostate cancer and its contribution to tumor cell

survival and stress response. Oncogene 30, 907–921 (2011).

125. Zhou, S.-J., Bi, T.-Q., Qin, C.-X., Yang, X.-Q. & Pang, K. Expression of NAMPT is associated

with breast invasive ductal carcinoma development and prognosis. Oncol. Lett. 15, 6648–

6654 (2018).

126. Van Beijnum, J. R. et al. Target validation for genomics using peptide-specific phage

antibodies: a study of five gene products overexpressed in colorectal cancer. Int. J. Cancer

101, 118–127 (2002).

127. Olesen, U. H. et al. Anticancer agent CHS-828 inhibits cellular synthesis of NAD. Biochem.

Biophys. Res. Commun. 367, 799–804 (2008).

128. Hasmann, M. & Schemainda, I. FK866, a highly specific noncompetitive inhibitor of

nicotinamide phosphoribosyltransferase, represents a novel mechanism for induction of

tumor cell apoptosis. Cancer Res. 63, 7436–7442 (2003).

129. Heske, C. M. Beyond Energy Metabolism: Exploiting the Additional Roles of NAMPT for

Cancer Therapy. Front. Oncol. 9, 1514 (2019).

130. Pan, L.-Z. et al. The NAD synthesizing enzyme nicotinamide mononucleotide

adenylyltransferase 2 (NMNAT-2) is a p53 downstream target. Cell Cycle vol. 13 1041–1048

(2014).

131. Cui, C. et al. Nicotinamide Mononucleotide Adenylyl Transferase 2: A Promising Diagnostic

and Therapeutic Target for Colorectal Cancer. Biomed Res. Int. 2016, 1804137 (2016).

132. Chalkiadaki, A. & Guarente, L. The multifaceted functions of sirtuins in cancer. Nat. Rev.

Cancer 15, 608–624 (2015).


133. Yu, D.-F. et al. Expression and clinical significance of Sirt1 in colorectal cancer. Oncol. Lett.

11, 1167–1172 (2016).

134. Li, X., Jiang, Z., Li, X. & Zhang, X. SIRT1 overexpression protects non-small cell lung cancer

cells against osteopontin-induced epithelial-mesenchymal transition by suppressing NF-κB

signaling. Onco. Targets. Ther. 11, 1157–1171 (2018).

135. Peck, B. et al. SIRT inhibitors induce cell death and p53 acetylation through targeting both

SIRT1 and SIRT2. Mol. Cancer Ther. 9, 844–855 (2010).

136. Lee, J. T. & Gu, W. SIRT1: Regulator of p53 Deacetylation. Genes & Cancer vol. 4 112–117

(2013).

137. Herranz, D. et al. SIRT1 promotes thyroid carcinogenesis driven by PTEN deficiency.

Oncogene 32, 4052–4056 (2013).

138. Marshall, G. M. et al. SIRT1 promotes N-Myc oncogenesis through a positive feedback loop

involving the effects of MKP3 and ERK on N-Myc protein stability. PLoS Genet. 7, e1002135

(2011).

139. Hanahan, D. & Weinberg, R. A. Hallmarks of Cancer: The Next Generation. Cell vol. 144

646–674 (2011).

140. Yi, M. et al. Advances and perspectives of PARP inhibitors. Exp. Hematol. Oncol. 8, 29

(2019).

141. Pilié, P. G., Gay, C. M., Byers, L. A., O’Connor, M. J. & Yap, T. A. PARP Inhibitors: Extending

Benefit Beyond BRCA-Mutant Cancers. Clinical Cancer Research vol. 25 3759–3771 (2019).

142. Faraoni, I. & Graziani, G. Role of BRCA Mutations in Cancer Treatment with Poly(ADP-

ribose) Polymerase (PARP) Inhibitors. Cancers 10, (2018).

143. Ashworth, A. & Lord, C. J. Synthetic lethal therapies for cancer: what’s next after PARP

inhibitors? Nat. Rev. Clin. Oncol. 15, 564–576 (2018).

144. Lord, C. J. & Ashworth, A. PARP inhibitors: Synthetic lethality in the clinic. Science 355,

1152–1158 (2017).
145. Murai, J. et al. Trapping of PARP1 and PARP2 by Clinical PARP Inhibitors. Cancer Res. 72,

5588–5599 (2012).

146. Pommier, Y., O’Connor, M. J. & de Bono, J. Laying a trap to kill cancer cells: PARP inhibitors

and their mechanisms of action. Sci. Transl. Med. 8, 362ps17 (2016).

147. Malavasi, F. et al. CD38 and chronic lymphocytic leukemia: a decade later. Blood vol. 118

3470–3478 (2011).

148. Gao, Y., Camacho, L. H. & Mehta, K. Retinoic acid-induced CD38 antigen promotes leukemia

cells attachment and interferon-γ/interleukin-1β-dependent apoptosis of endothelial cells:

Implications in the etiology of retinoic acid syndrome. Leukemia Research vol. 31 455–463

(2007).

149. Leo, R. et al. Multiparameter analyses of normal and malignant human plasma cells:

CD38++, CD56+, CD54+, cIg+ is the common phenotype of myeloma cells. Ann. Hematol.

64, 132–139 (1992).

150. Lin, P., Owens, R., Tricot, G. & Wilson, C. S. Flow Cytometric Immunophenotypic Analysis

of 306 Cases of Multiple Myeloma. American Journal of Clinical Pathology vol. 121 482–488

(2004).

151. Bu, X. et al. CD38 knockout suppresses tumorigenesis in mice and clonogenic growth of

human lung cancer cells. Carcinogenesis 39, 242–251 (2018).

152. Morandi, F. et al. CD38: A Target for Immunotherapeutic Approaches in Multiple Myeloma.

Front. Immunol. 9, 2722 (2018).

153. Shallis, R. M., Terry, C. M. & Lim, S. H. The multi-faceted potential of CD38 antibody targeting

in multiple myeloma. Cancer Immunol. Immunother. 66, 697–703 (2017).

154. Petrucci, M. T. & Vozella, F. The Anti-CD38 Antibody Therapy in Multiple Myeloma. Cells 8,

(2019).

155. Chatterjee, S. et al. CD38-NADAxis Regulates Immunotherapeutic Anti-Tumor T Cell

Response. Cell Metab. 27, 85–100.e8 (2018).


156. Verma, V. et al. PD-1 blockade in subprimed CD8 cells induces dysfunctional PD-1CD38

cells and anti-PD-1 resistance. Nat. Immunol. 20, 1231–1243 (2019).

157. Chen, L. et al. CD38-Mediated Immunosuppression as a Mechanism of Tumor Cell Escape

from PD-1/PD-L1 Blockade. Cancer Discov. 8, 1156–1175 (2018).

158. Ortolan, E. et al. Functional role and prognostic significance of CD157 in ovarian carcinoma.

J. Natl. Cancer Inst. 102, 1160–1177 (2010).

159. Krupka, C. et al. Targeting CD157 in AML using a novel, Fc-engineered antibody construct.

Oncotarget 8, 35707–35717 (2017).

160. Lee, H. C. Cyclic ADP-Ribose and NAADP: Structures, Metabolism, and Functions. (Springer

Science & Business Media, 2002).

161. Yoshino, J. & Imai, S.-I. Accurate Measurement of Nicotinamide Adenine Dinucleotide (NAD

) with High-Performance Liquid Chromatography. Sirtuins 203–215 (2013) doi:10.1007/978-

1-62703-637-5_14.

162. Formentini, L., Moroni, F. & Chiarugi, A. Detection and pharmacological modulation of

nicotinamide mononucleotide (NMN) in vitro and in vivo. Biochemical Pharmacology vol. 77

1612–1620 (2009).

163. Trammell, S. A. & Brenner, C. Targeted, LCMS-based Metabolomics for Quantitative

Measurement of NAD(+) Metabolites. Comput. Struct. Biotechnol. J. 4, e201301012 (2013).

164. Yang, H. et al. Nutrient-sensitive mitochondrial NAD+ levels dictate cell survival. Cell 130,

1095–1107 (2007).

165. Gross, C. J. & Henderson, L. M. Digestion and Absorption of NAD by the Small Intestine of

the Rat. The Journal of Nutrition vol. 113 412–420 (1983).

166. Grozio, A. et al. Slc12a8 is a nicotinamide mononucleotide transporter. Nat Metab 1, 47–57

(2019).

167. Liu, L. et al. Quantitative Analysis of NAD Synthesis-Breakdown Fluxes. Cell Metab. 27,

1067–1080.e5 (2018).
168. Frederick, D. W. et al. Loss of NAD Homeostasis Leads to Progressive and Reversible

Degeneration of Skeletal Muscle. Cell Metab. 24, 269–282 (2016).

169. Scheibye-Knudsen, M. et al. A high-fat diet and NAD(+) activate Sirt1 to rescue premature

aging in cockayne syndrome. Cell Metab. 20, 840–855 (2014).

170. Fang, E. F. et al. NAD Replenishment Improves Lifespan and Healthspan in Ataxia

Telangiectasia Models via Mitophagy and DNA Repair. Cell Metab. 24, 566–581 (2016).

171. Zhang, T. & Lee Kraus, W. SIRT1-dependent regulation of chromatin and transcription:

Linking NAD metabolism and signaling to the control of cellular functions. Biochimica et

Biophysica Acta (BBA) - Proteins and Proteomics vol. 1804 1666–1675 (2010).

172. Fang, E. F. Mitophagy and NAD inhibit Alzheimer disease. Autophagy vol. 15 1112–1114

(2019).

173. Mouchiroud, L. et al. The NAD(+)/Sirtuin Pathway Modulates Longevity through Activation of

Mitochondrial UPR and FOXO Signaling. Cell 154, 430–441 (2013).

174. Cantó, C. et al. AMPK regulates energy expenditure by modulating NAD metabolism and

SIRT1 activity. Nature vol. 458 1056–1060 (2009).

175. Fulco, M. et al. Glucose Restriction Inhibits Skeletal Myoblast Differentiation by Activating

SIRT1 through AMPK-Mediated Regulation of Nampt. Developmental Cell vol. 14 661–673

(2008).

176. Price, N. L. et al. SIRT1 is required for AMPK activation and the beneficial effects of

resveratrol on mitochondrial function. Cell Metab. 15, 675–690 (2012).

177. Covarrubias, A. J. et al. Senescent cells promote tissue NAD decline during ageing via the

activation of CD38 macrophages. Nature Metabolism vol. 2 1265–1283 (2020).

178. Chini, C. C. S. et al. CD38 ecto-enzyme in immune cells is induced during aging and

regulates NAD and NMN levels. Nat Metab 2, 1284–1304 (2020).

179. Igarashi, M. et al. NAD supplementation rejuvenates aged gut adult stem cells. Aging Cell

18, e12935 (2019).


180. Vannini, N. et al. The NAD-Booster Nicotinamide Riboside Potently Stimulates

Hematopoiesis through Increased Mitochondrial Clearance. Cell Stem Cell 24, 405–418.e7

(2019).

181. Rimmelé, P. et al. Aging-like Phenotype and Defective Lineage Specification in SIRT1-

Deleted Hematopoietic Stem and Progenitor Cells. Stem Cell Reports vol. 3 44–59 (2014).

182. Lee, H. J., Hong, Y.-S., Jun, W. & Yang, S. J. Nicotinamide Riboside Ameliorates Hepatic

Metaflammation by Modulating NLRP3 Inflammasome in a Rodent Model of Type 2 Diabetes.

J. Med. Food 18, 1207–1213 (2015).

You might also like