Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Niacin: James B. Kirkland, Mirella L. Meyer-Ficca

Download as pdf or txt
Download as pdf or txt
You are on page 1of 67

CHAPTER THREE

Niacin
James B. Kirkland*, Mirella L. Meyer-Ficca†,1
*Department of Human Health and Nutritional Sciences, University of Guelph, Guelph, ON, Canada

Utah State University, Logan, UT, United States
1
Corresponding author: e-mail address: mirella.meyer@usu.edu

Contents
1. Introduction 84
2. Role of Dietary Niacin in Disease Origin and Progression 84
2.1 Historic Niacin-Deficiency Syndrome: Pellagra 84
2.2 Current Risks and Causes of Niacin Deficiency and Pellagra 86
2.3 Niacin and Vitamin B3, Assessment of Niacin Status, and Recommended
Daily Allowance 88
2.4 Niacin Food Sources 91
2.5 Risk Factors Predisposing to Niacin Deficiency 91
3. NAD Metabolism 93
3.1 NAD Synthesis 93
3.2 NAD Recycling 97
3.3 NADPH Synthesis 97
3.4 NAD and NADP Utilization 97
3.5 NAD as Substrate 105
3.6 Turnover of PAR and MAR Modifications 117
4. NAD: Relevance to Health and Disease 118
4.1 NAD Decline During Aging and Age-Related Disorders 119
4.2 DNA Damage Accumulation 121
4.3 Cell Death/Cancer 122
4.4 Mitochondrial Dysfunction 122
4.5 Neurological Disorders 124
4.6 Skin Health and Skin Cancer Prevention 127
4.7 Cardiovascular Disorders and Niacin in Clinical Trials 128
5. Summary 132
References 132

Abstract
Nicotinic acid and nicotinamide, collectively referred to as niacin, are nutritional precur-
sors of the bioactive molecules nicotinamide adenine dinucleotide (NAD) and nicotin-
amide adenine dinucleotide phosphate (NADP). NAD and NADP are important
cofactors for most cellular redox reactions, and as such are essential to maintain cellular
metabolism and respiration. NAD also serves as a cosubstrate for a large number of ADP-
ribosylation enzymes with varied functions. Among the NAD-consuming enzymes

Advances in Food and Nutrition Research, Volume 83 # 2018 Elsevier Inc. 83


ISSN 1043-4526 All rights reserved.
https://doi.org/10.1016/bs.afnr.2017.11.003
84 James B. Kirkland and Mirella L. Meyer-Ficca

identified to date are important genetic and epigenetic regulators, e.g., poly(ADP-
ribose)polymerases and sirtuins. There is rapidly growing knowledge of the close
connection between dietary niacin intake, NAD(P) availability, and the activity of
NAD(P)-dependent epigenetic regulator enzymes. It points to an exciting role of dietary
niacin intake as a central regulator of physiological processes, e.g., maintenance of
genetic stability, and of epigenetic control mechanisms modulating metabolism and
aging. Insight into the role of niacin and various NAD-related diseases ranging from
cancer, aging, and metabolic diseases to cardiovascular problems has shifted our view
of niacin as a vitamin to current views that explore its potential as a therapeutic.

1. INTRODUCTION
Niacin (nicotinic acid or vitamin B3) is a functional group present in
the coenzymes nicotinamide adenine dinucleotide (NAD) and nicotinamide
adenine dinucleotide phosphate (NADP), which are essential for oxidative
processes. Niacin was first identified in 1937 as the dietary factor that
prevented black tongue disease in dogs, the animal model for pellagra
(Koehn & Elvehjem, 1937). Pellagra is a dietary-deficiency disease whose
study facilitated the identification of niacin as the pellagra-preventing factor,
the discovery of niacin as the essential vitamin required for adequate NAD
synthesis in humans, and the identification of multiple roles of coenzyme
NAD(P) in metabolism.

2. ROLE OF DIETARY NIACIN IN DISEASE ORIGIN AND


PROGRESSION
2.1 Historic Niacin-Deficiency Syndrome: Pellagra
The disease pellagra was first described in 1763 by the Spanish physician
Casal, soon after maize (corn) became a major part of the European diet.
In 1771, Francesco Frapolli, an Italian physician, coined the term
“pellagra” (literally: sour or rough skin) to describe the crusty and inflamed
skin characteristic in human patients suffering from dietary niacin wasting
deficiency.
In 1870, frequent pellagra outbreaks among poor populations in
southern and central Europe had followed the introduction of maize.
At the end of the 19th and the beginning of the 20th centuries, pellagra
reached epidemic proportions among the population of the southern
United States and became the most severe deficiency disease in the history
of the United States, with more than 120,000 deaths by the 1920s
Niacin 85

(Etheridge, 1972; Rajakumar, 2000). Joseph Goldberger, working for the


United States Public Health Service in the 1920s, discovered that milk
and meat could provide a pellagra-preventing (PdP) factor, which was
absent in corn.
In 1937, Koehn and Elvehjem used liver extracts to isolate the dietary
PdP factor that prevented black tongue disease in a dog model of pellagra
and subsequently identified it as nicotinic acid (Elvehjem, Madden,
Strong, & Woolley, 1937; Koehn & Elvehjem, 1937). This discovery laid
the foundation for more studies that confirmed lack of nicotinic acid as
the underlying cause of pellagra in humans, and the rapid implementation
of nicotinic acid and nicotinamide supplementation as the efficient
treatment.
It was initially puzzling that Latin American populations had been spared
from pellagra, despite using corn as their main staple food for centuries. This
riddle was solved when differences in the traditional food processing
methods were discovered. Traditional Latin American food processing prac-
tices involved presoaking corn in alkaline “lime-water” prior to cooking,
which released the niacin from binding to glycoproteins, increasing its
bioavailability.
In 1942, the US government mandated a niacin enrichment program for
flour, and similar programs in other developed countries have diminished
the occurrence of clinical pellagra to sporadic cases associated with malnu-
trition (e.g., neglect, abuse, or anorexia nervosa), alcoholism, iatrogenic sit-
uations (e.g., cancer treatments), malabsorption, and prolonged diarrhea.
In the 1960s and 1970s, pellagra still was considered a public health con-
cern in developing African and Asian countries, where seasonally between
15% and 50% of the population presented with skin manifestations of pel-
lagra (Aykroyd, 1971; Barrett-Connor, 1967; Gopalan, 2009). Pellagra out-
breaks still occur frequently among refugee populations, where they
continue to be of ongoing concern.
Pellagra became known as the disease of the “4 D’s,” referring to three
commonly seen components (diarrhea, dermatitis, and dementia) of the syn-
drome culminating in death (the 4th D) (Hegyi, Schwartz, & Hegyi, 2004).
The first D, diarrhea, the gastrointestinal presentation of pellagra, results
from intestinal inflammation, when the normally metabolically active and
rapidly dividing mucosal tissue is affected. Patients appear nauseous and have
poor appetite and abdominal pain. Anorexia and malabsorptive diarrhea
exacerbate a state of malnutrition.
The second D, dermatitis, refers to extensive skin lesions. After an initial
erythematous phase resembling sunburn with blisters, the skin becomes
86 James B. Kirkland and Mirella L. Meyer-Ficca

thickened and hyperpigmented when dermatitis progresses. Typical for


pellagra is a symmetrical and bilateral distribution of affected regions in
sun-exposed skin, e.g., dorsal parts of hands, feet, cheeks, forehead, and
sun-exposed regions of the neck, sometimes referred to as “Casal’s necklace.”
The third D refers to frequently observed neurological symptoms, e.g.,
muscle weakness, and the frequent deterioration of mental health observed
in pellagra patients, progressing from lethargy, apathy, and depression to
anxiety and irritability. While these symptoms are not surprising during
severe malnutrition, many patients with pellagrous dementia experience
visual and auditory hallucinations and display paranoid and aggressive behav-
iors, which can disappear within hours of niacin supplementation,
suggesting impairment of brain signaling rather than structural degeneration.
The final D, death, indicates the high mortality rate observed in severe
pellagra cases when depletion of the niacin-based coenzyme NAD prevents
maintenance of vital body functions.

2.2 Current Risks and Causes of Niacin Deficiency and Pellagra


In addition to a diet deficient in niacin or its precursor tryptophan, the com-
mon causes of pellagra, other mechanisms can contribute to suboptimal nia-
cin status. Primary causes of pellagra include dependence on a corn-based
diet, or general malnutrition (associated with poverty, neglect, abuse, and
famine, as well as anorexia nervosa) (Jagielska, Tomaszewicz-Libudzic, &
Brzozowska, 2007). Some secondary causes include chronic alcoholism
(Nogueira, Duarte, Magina, & Azevedo, 2009) and general malabsorptive
states such as prolonged diarrhea. Alcohol-associated pellagra cases continue
to be reported in countries with niacin fortification (Li et al., 2016; Terada,
Kinoshita, Taguchi, & Tokuda, 2015). In scientific and medical reports of
pellagra cases, alcoholism is the most frequently identified etiology (35%)
compared to medications (56%), inadequate intake and malabsorption
(16% and 13%, respectively), metabolic disorders (8%), and infrequent
unknown etiologies (Goudarzi et al., 2016). When diarrhea sets in as part
of the progression of pellagra, it tends to cause a vicious cycle of worsening
niacin status.
In addition, other secondary factors induce pellagra-like symptoms and
contribute to impaired NAD synthesis:
2 Diets rich in leucine, like those based on Sorghum, can lead to pellagra
due to altered tryptophan metabolism in the presence of excessive leu-
cine (Beretich, 2005; Gopalan, 1969, 2009).
Niacin 87

2 Hartnup disease is an autosomal recessive mutation in the gene for the


neutral amino acid transporter SLC6A19, leading to depletion of tryp-
tophan and increased susceptibility for development of pellagra (Br€ oer,
Cavanaugh, & Rasko, 2005; Kraut & Sachs, 2005; Nozaki et al., 2001).
Administration of niacin reverts the pellagra symptoms and highlights
tryptophan’s importance as complementary NAD precursor.
2 Carcinoid tumors can cause secondary tryptophan degradation due to
excessive serotonin production. As tryptophan is a precursor for both,
NAD and serotonin synthesis, an increased tryptophan demand via an
increase in serotonin synthesis depletes tryptophan that is available for
NAD synthesis resulting in niacin deficiency and occasional pellagra-like
symptoms (Bouma et al., 2016; Shah et al., 2005).
2 Infections with human immunodeficiency virus (HIV) have been asso-
ciated with decreased plasma tryptophan levels and a pellagra-like state
that responded to nicotinamide supplementation (Murray, Langan, &
MacGregor, 2001).
2 Additional diseases, e.g., liver cirrhosis, diabetes mellitus, and prolonged
fevers, have been associated with niacin deficiency and pellagra.
2 Increasing age correlates with dropping NAD+ levels (Braidy et al.,
2011). Because nicotinamide and NAD+ levels are currently moving
to the limelight as players determining longevity (Li, Chong, &
Maiese, 2006; Verdin, 2015), current research aims to augment intracel-
lular NAD+ to prevent aging-associated disorders and decline
(Srivastava, 2016).
2 Women have about twice the risk of niacin deficiency compared to men
(Miller, 1978). A study of sex-specific effects of tryptophan to niacin
conversion in rats found a sex hormone-dependent decrease in the effi-
ciency of tryptophan to niacin and NAD synthesis in females. This sex-
specific difference puts women, and in particular pregnant females, at an
increased risk for niacin deficiency and pellagra (Shibata & Toda, 1997).
Indeed, a recent human study measured blood vitamin concentrations in
pregnant women during the trimester of pregnancy. The women were
healthy and eating a good diet supplemented with vitamins. Despite
hypervitaminemic levels of folate, biotin, pantothene, and riboflavin,
blood levels of vitamin A, niacin, thiamine, and vitamin B12 were defi-
cient. The niacin hypovitaminemia was especially pronounced. A high
percentage of pregnant women were deficient already in the first trimes-
ter, and this deficiency worsened in the later trimesters (Baker,
DeAngelis, Holland, Gittens-Williams, & Barrett, 2002).
88 James B. Kirkland and Mirella L. Meyer-Ficca

2.3 Niacin and Vitamin B3, Assessment of Niacin Status, and


Recommended Daily Allowance
2.3.1 Vitamin B3 Structures
Pellagra is the manifestation of a deficiency in NAD. Molecules that serve as
dietary precursors for NAD, other than tryptophan, constitute the vitamin
B3 family. Fig. 1 gives an overview of currently known vitamin B3 mole-
cules. The first identified “pellagra-preventing factor,” i.e., vitamin B3, was
nicotinic acid or pyridine-3-carboxylic acid, subsequently referred to as nia-
cin (nicotinic acid vitamin), and this is the main source of B3 found in plant-
based foods. Nicotinamide or pyridine-3-carboxamide, sometime called
niacinamide, also serves as nutritional NAD precursor and prevents pellagra
through a similar, albeit not completely identical pathway. Nicotinamide is
formed by the cleavage of NAD+ and is the main circulating form of B3 in
blood. Despite this original distinction in nomenclature between niacin and
niacinamide, today the term niacin often indiscriminately refers to both, nic-
otinic acid or nicotinamide, or a combination of both. In 2004, Bieganowsi
and Brenner discovered that nicotinamide ribose, a molecule present, e.g., in
milk, can similarly be converted to NAD and thus classifies as the newest
discovered member of the vitamin B3 family (Bieganowski &
Brenner, 2004).

2.3.2 Measurement of Niacin Status


Since the availability of dietary niacin in humans is directly reflected by
NAD content in the body, biochemical determination of blood or tissue
NAD is utilized as a measure of niacin status (Jacobson & Jacobson, 1997;
Shah et al., 2005). A study determining blood NAD content in a metabolic
ward demonstrated that the NAD content in red blood cells can be used as a
marker for niacin status in humans. Blood NAD content directly mirrored
low niacin diet or niacin supplementation in the form of either decreased or
increased blood NAD levels (Fu, Swendseid, Jacob, & McKee, 1989;
Jackson, Rawling, Roebuck, & Kirkland, 1995; Jacobson, Shieh, &
Huang, 1999). Since blood NADP content remains relatively stable even
with a niacin poor diet that causes NAD to drop (Tang, Sham, Hui, &
Kirkland, 2008), the ratio of NAD to NADP (called niacin number) is some-
times used as a measure of niacin status. Interestingly, even in developed
countries with niacin-fortified flour, blood niacin numbers vary to a surpris-
ing degree, and 15%–20% of the population were found niacin deficient
(Jacobson, 1993).
Alternatively, quantification of urinary metabolites, i.e.,
N-methylnicotinamide and 2-pyridone, can be measured to determine
Niacin 89

H H
N

O
O O
+
N
H H
O N
O
H H O

N N
O O H
Nicotinic acid Nicotinamide H
niacin “niacinamide”
Nicotinamide riboside
+ PRPP, ATP
+P
RP

P
AT
P
,A

2
TP

+
,Q

H N
Nicotinamide adenine N
dinucleotide (NAD+) N
H N
N
H N O H
O
O

+
N O O O H
O
O P H
O
HO O P O
O O−
H

Fig. 1 Vitamin B3 molecules (nicotinic acid, nicotinamide, and nicotinamide riboside)


are dietary precursors that support the formation of nicotinamide adenine dinucleotide
(NAD). Cofactors for the synthesis of NAD from the respective precursor are indicated as
follows: ATP, adenosine triphosphate; PRPP, 5-phosphoribosyl-ribose-1-pyrophosphate;
Q, glutamate.

recent dietary niacin intake and niacin status at an individual and population
level, with the caveat that 2-pyridone more likely indicates adequate protein
intake than niacin status.

2.3.3 Dietary Need and Recommended Daily Intake


Dietary need for NAD synthesis is usually expressed as of “niacin”
requirements, i.e., the direct precursors for NAD synthesis taken up as vita-
min B3 molecules. A small percentage of NAD is generated from the
90 James B. Kirkland and Mirella L. Meyer-Ficca

essential amino acid tryptophan via the de novo NAD synthesis. The effi-
ciency of the tryptophan to NAD metabolism is estimated to require
60 mg tryptophan for the generation of 1 mg niacin, so that a 60 mg trypto-
phan intake counts as 1 niacin equivalent (NE) (Horwitt, Harper, &
Henderson, 1981).
In the United States, the Food and Drug Administration regulates vita-
mins as “dietary supplements” and requires vitamin manufacturers to follow
current GMP regulations. To prevent and correct dietary insufficiencies in
certain food products, most flour and grain products produced as “enriched
products” must contain specified levels of thiamin, niacin, riboflavin, iron,
and folic acids. According to 21 CFR137.165 (fda.gov; revised as of April 1,
2016), each pound of flour must contain 24 mg of niacin.
The recommended daily allowance (RDA) is the amount of nutrient
or vitamin that is enough to meet the nutrient requirements of nearly all
(97%–98%) healthy people. RDA for niacin in adults is 16 mg/day of NE
for men and 14 mg of NE for women according to the FOA/WHO, status
1996 (see Table 1). For children, these values are adjusted to age groups and
gradually increase from 6 NEs daily in toddlers to 12 in teenagers, until they
reach the adult requirements. Recent studies estimate the median daily
intake of preformed dietary niacin in the United States to be 28 mg for
men and 18 mg for women. In Canada, median daily intake was approxi-
mately 41 mg in men and 28 mg in women (Institute of Medicine, 1998).
Chronic high oral doses of nicotinic acid can induce hepatotoxicity and der-
matological problems in the form of skin flushing, so that the US Food and
Nutrition Board has recommended a 35 mg tolerable upper limit (UL) for
daily intake (Institute of Medicine, 1998).

2.3.4 Factors That Influence Dietary Niacin Requirement


One important factor influencing the dietary niacin requirements is the bio-
availability of the niacin that is present in available food. In mature cereal
grains, the majority of niacin is bound and only about 30% is bioavailable.
Beans and liver are among the foods that contain niacin in the free form. As
mentioned, tryptophan as an NAD precursor in humans is thought to be
converted to NAD at an efficiency of 60:1 to 70:1 (Horwitt et al., 1981).
This conversion rate has been found to vary widely (30%) between individ-
uals and tends to increase with increased tryptophan consumption, likely
because more tryptophan is available for conversion to NAD once trypto-
phan demand for protein synthesis is met (Patterson, Brown, Linkswiler, &
Harper, 1980). Factors that reduce this conversion efficiency increase the
Niacin 91

Table 1 Recommended Dietary Allowance (RDA) of Niacin Intake for Humans at


Different Point in the Life Span
Group Age (Years) Recommended Intake (NE/Day)
Infants (AI) 0–0.5 2
0.5–1 4
Children (RDA) 1–3 6
4–8 8
9–13 12
Adolescents and adults (RDA)
Males >14 16
Females >14 14
Pregnant 18
Lactating 17
AI, adequate intake, recommendation when there is not enough evidence to develop an RDA; NE, nia-
cin equivalent, defined as mg niacin + 1/60 mg tryptophan; RDA, recommended daily allowance, aver-
age daily level of intake that is thought enough to meet the nutrient requirements of nearly all (97%–98%)
healthy people.
USDA Agricultural Research Services, Dietary Reference Intakes, Information Retrieved on April 7,
2017.

requirement for niacin. Examples in addition to low tryptophan intake are


carcinoid syndrome, in which increased serotonin synthesis depletes trypto-
phan, and Hartnup disease, which is characterized by deficient tryptophan
uptake, and long-term treatment with the drug Isoniazid, which interferes
indirectly with the tryptophan to NAD pathway.

2.4 Niacin Food Sources


Plants, bacteria, and yeast produce niacin and nicotinamide. From there, the
vitamin molecules are maintained in the food chain. Niacin content varies in
different food sources (see Table 2), and various food groups are suitable
sources for dietary niacin. Fish, meat, milk, peanuts, and enriched flour
products are good niacin sources for human consumption.

2.5 Risk Factors Predisposing to Niacin Deficiency


The major risk factor for niacin deficiency is the ongoing consumption of a
diet that relies mostly on nonfortified refined grains and grain products, as
well as a diet with little variety that is low in animal or dairy products
and legumes. Maize- or Sorghum-based diets are examples of such diets
92 James B. Kirkland and Mirella L. Meyer-Ficca

Table 2 Preformed Niacin Content in Food


Food Niacin Content (Values in mg per 100 g or 100 mL)
Meats and Fish Fish (tuna, skipjack) 12–19
Fish (salmon, mahimahi, swordfish) 7–9
Poultry (chicken, turkey) 9–12
Pork 6–11
Beef and veal 7–9
Game meat, bison 7–11
Dairy and Eggs Milk (liquid, whole—nonfat) 0.09–0.1
Milk (dry, nonfat) 0.7–1
Yoghurt (various kinds) 0.1–0.2
Ice cream (various kinds) 0.07–1.5
Cheese 0.03–1.2
Egg 0.075–0.2
Grains, Cereals, ready to eat 0.2–90
Cereals
Rice, uncooked 5
Pasta and noodles, enriched, dry 6–8.4
Flour (wheat, barley, rice, buckwheat, white, 6–7.5
enriched)
Cornmeal (enriched) 4.9–5.2
Corn flour, cornmeal, corn grain 2.6–3.6
Vegetables Mushrooms 3–14
Tomato products (canned, paste) 0.9–3
Beans 2–3
Corn 1.1–1.8
Potatoes 1.3–1.9
Mustard greens, spinach, cabbage, cauliflower 0.2–0.3
Yeast extract spread 128
Legumes Peanuts (raw and roasted nuts, butter) 12–15
Soy products (beans, meal, flour, sauce) 1–3.2
Soymilk 0.2
Beans (kidney, garbanzo, mung, black, adzuki, lima) 0.4–0.7
Niacin 93

Table 2 Preformed Niacin Content in Food—cont’d


Food Niacin Content (Values in mg per 100 g or 100 mL)
Nuts Almonds (raw, roasted, butter, paste) 1.4–3.7
Cashew (raw and roasted nuts, butter) 1–1.7
Coconut (milk, meat, cream) 0.3–0.9
Sunflower seed kernels (dried, roasted, butter) 4–8
Sesame seeds (meal, flour, paste, butter) 4.5–13.4
Fruit Physalis (groundcherries, fresh poha) 2.8
Avocados (raw) 1.9
Oranges 0.4
Strawberries 0.4
Apples 0.08
Baked goods Leavening agents (active dried yeast) 40.2
Bagels 1.8–4.6
Bread 2–6
Waffles 2–7
USDA Agricultural Research Services, Food Composition Database (version from March 29, 2017).

low in bioavailable niacin. Sorghum (millet, jowar) contains suitable amounts


of tryptophan, but excessive amounts of leucine in Sorghum interfere with
utilization of tryptophan for niacin synthesis.
Certain population groups are at risk for niacin deficiency even in the
context of supplemented food products, i.e., secondary niacin deficiency.
This is usually due to either increased niacin demand with increased meta-
bolic NAD consumption, e.g., as seen in elderly people and pregnant
women, or cancer patients undergoing treatments that induced DNA dam-
age, such as radiation therapy or exposure to DNA-damaging drugs.

3. NAD METABOLISM
3.1 NAD Synthesis
NAD can be synthesized from nicotinic acid (NA), nicotinamide (NAM),
nicotinamide riboside (NR) (all members of the vitamin B3 molecule
group), and tryptophan.
NAD is synthesized from NA by the Preiss–Handler pathway. It comprises
three individual metabolic conversions (Fig. 2). In a first step, NA is
Vitamin B3 Tryptophan
O

O
Nicotinamide Nicotinamide Nicotinic acid N N
H
H H
“Niacinamide” Riboside Niacin H

O
H N H
O TDO
O IDO1/IDO2
H O
N +
N H
H H
O N H O
N O
N
H O

H
O O H O N-Formyl-
O HNH H
kynurenine

NAMPT1 NRK1 NAPRT1 AFMID


NAD used as substrate for enzymatic reactions,

NRK2 Phosphoribosyl
Phosphoribosyl KMO
pyrophosphate pyrophosphate
ATP
KYNU
PPi PPi
NAD salvage pathway

PPi
HAAO
H
O O
O

O ACMSD
N H
HH H

ACMS
Acetyl CoA
Phospho- O
CO2
H ribosyl pyro- Nonenzymatic,
O N +
H
O
H phosphate spontaneous
Nicotinamide –
O
O
PPi H
O P O H
O
mononucleotide O
O
O
N+
H +
O
(NMN) H
O N
O


O
Nicotinic acid QPRT N
H
O P O

H O
O
H mononucleotide Quinolinic
O O O
(NAMN) H acid
ATP ATP
NMNAT1–3 NMNAT1–3 PPi
PPi

H N
Nicotinamide adenosine N
H N

dinucleotide (NAD+) N
H
N

N
H N
O N
O
H
N O
N
H NADS H
O
O O
O
N+ O O H
O O
+
O H O–
N
O
O
O Glutamate Glutamine P
O
H
H O P O
O
O P
O
H + AMP + ATP O O
H O P O H
O
+ PPi
O O–
H
Nicotinic acid adenine dinucleotide
(NAAD)

Fig. 2 NAD synthesis pathways present in mammals. Blue arrows illustrate the Preiss–
Handler pathway of NAD synthesis from the nicotinic acid conversion, and the NAD sal-
vage pathway comprises NAD synthesis from nicotinamide (green arrows). NAD from
tryptophan occurs via the de novo synthesis pathway (brown arrows). Abbreviations:
ACMSD, ACMS decarboxylase [EC:4.1.1.45]; ACMS, alpha-amino-beta-carboxy-
muconate-semialdehyde; AFMID, N-formylkynurenine formamidase [EC:3.5.1.9]; HAAO,
3-hydroxyanthranilate 3,4-dioxygenase [EC:1.13.11.6]; IDO1, indoleamine 2,3-
dioxygenase 1 [EC:1.13.11.52]; IDO2, indoleamine 2,3-dioxygenase 2 [EC:1.13.11.52];
KMO, kynurenine 3-monooxygenase [EC:1.14.13.9]; KYNU, kynureninase [EC:3.7.1.3];
NAAD, nicotinic acid adenine dinucleotide; NAD, nicotinamide adenine dinucleotide;
NADS, NAD synthetase [EC:6.3.5.1]; NAMN, nicotinic acid mononucleotide; NAMPT, nic-
otinamide phosphoribosyl transferase [EC:2.4.2.12]; NAPRT, nicotinic acid phospho-
ribosyl transferase [EC:6.3.4.21]; NMN, nicotinamide mononucleotide; NMNAT,
nicotinamide mononucleotide adenylyltransferase [EC:2.7.7.1 2.7.7.18]; NR, nicotin-
amide riboside; NRK1, NR kinase 1 [EC:2.7.1.22 2.7.1.173]; NRK2, NR kinase 2
[EC:2.7.1.22 2.7.1.173]; QPRT, quinolinate phosphoribosyl transferase [EC:2.4.2.19];
TDO, tryptophan 2,3-dioxygenase [EC:1.13.11.11].
Niacin 95

converted to nicotinic acid mononucleotide (NAMN) by the enzyme nic-


otinic acid phosphoribosyl transferase (NAPT, encoded for by the gene
NPT1) and use of 5-phosphate-α-D-ribose 1-diphosphate as cosubstrate.
NAMN is then converted to nicotinic acid adenine dinucleotide
(NAAD) by the enzyme NMNAT (nicotinamide mononucleotide
adenylyltransferase) using ATP as cosubstrate. Three isoforms of NMNAT
enzymes with distinct subcellular localizations have been described in
humans, which permits differentially regulated subcellular NAD concentra-
tions (Emanuelli et al., 2001; Raffaelli et al., 2002; Zhang et al., 2003). As a
final step toward NAD synthesis, the NA moiety is amidated by the enzyme
NADS (NAD synthetase), which transfers an amino group from glutamine
and simultaneously forms glutamate in an ATP-consuming step.
The synthesis of NAD from nicotinamide follows the salvage pathway.
The name reflects the role of this pathway in the reutilization of nicotin-
amide released by ADP-ribosylases and NAD glycohydrolases in the body.
The enzyme nicotinamide phosphoribosyl transferase (NAMPT) performs
the initial step from nicotinamide to nicotinamide mononucleotide
(NMN). This step is rate limiting for local NAD synthesis, and NAMPT
activity is considered an important regulator of several downstream
NAD-consuming enzymes, e.g., sirtuins (Revollo, Grimm, & Imai, 2004,
2007). Making more NAD available is presumed to be the mechanism by
which NAMPT overexpression delays senescence in human cells (van der
Veer et al., 2007). Additional relevance for NAMPT comes from observa-
tions that the expression of NAMPT is regulated by the circadian rhythm
machinery. This makes it a direct mediator between circadian rhythms
and NAD salvage pathway (Nakahata, Sahar, Astarita, Kaluzova, &
Sassone-Corsi, 2009; Ramsey et al., 2009). The final step toward NAD syn-
thesis is performed by the NMNAT enzymes that condense the adenylyl
moiety to NMN in a manner comparable to the NAMN to NAD step dis-
cussed earlier.
The third molecule of the vitamin B3 group, nicotinamide riboside,
comprises the phosphorylation of NR by the enzyme NR kinase (NRK)
to form NMN, which subsequently serves as a substrate for NAD synthesis
via NMNAT adenylyl transfer, as discussed. Two isoforms for NRK have
been described (NRK1 and NRK2), and this pathway appears to be highly
conserved between species, e.g., it was found in yeast and humans (Belenky
et al., 2007; Bieganowski & Brenner, 2004).
An additional important pathway of NAD synthesis involves the
conversion of tryptophan to NAD via the de novo NAD synthesis pathway
96 James B. Kirkland and Mirella L. Meyer-Ficca

(right section in Fig. 2). This pathway involves seven biochemical reactions
in the kynurenine pathway, which transforms tryptophan to NAMN. NAMN
is transformed to NAAD and finally to NAD by NMNAT/NADS activities,
using the identical metabolic transformation that synthesizes NAD from the
NAMN generated by the NA metabolisms in the Preiss–Handler pathway.
The first, and rate-limiting step, of the de novo synthesis pathway
involves the conversions of tryptophan to n-formylkynurenine. This step
can be catalyzed by each of three enzymes: indoleamine 2,3-dioxygenase
1 (IDO1), indoleamine 2,3-dioxygenase 2 (IDO2), or tryptophan 2,3-
dioxygenase (TDO). These three enzymes have partially overlapping
expression patterns and potentially distinct functions (Ball, Jusof,
Bakmiwewa, Hunt, & Yuasa, 2014). TDO is the major enzyme relevant
for de novo synthesis in liver, while IDOs have highest activity in lung, small
intestine, and placenta (Kudo & Boyd, 2000; Yamazaki, Kuroiwa,
Takikawa, & Kido, 1985).
The four subsequent enzymatic reactions result in the formation of
alpha-amino-beta-carboxy-muconate-semialdehyde (ACMS). ACMS has
two distinct fates that influence niacin status; either it is decarboxylated
by the enzyme ACMS decarboxylase to alpha-amino-muconate-
semialdehyde (AMS), or it undergoes spontaneous enzyme-independent
cyclization and forms quinolinic acid. Quinolinic acid subsequently serves
as a precursor for NAD synthesis. AMS on the other hand serves as a sub-
strate for a pathway, leading to the formation of acetyl CoA. The efficiency
of NAD synthesis from tryptophan therefore hinges on the accumulation of
ACMS. High activity of ACMSD can enzymatically deplete this interme-
diate from the NAD-directed pathway and prevent efficient tryptophan
to NAD conversion, while low levels of ACMSD activity allow enough
ACMS accumulation to form quinolinic acid (Fukuwatari, Sugimoto, &
Shibata, 2002; Fukuwatari, Ohta, Kimtjra, Sasaki, & Shibata, 2004).
Quinolinate phosphoribosyl transferase (QPRT) condenses quinolinic acid
with 5-phospho-alpha-D-ribose 1-diphosphate to form NA mononucleo-
tide (NAMN), which is further metabolized to NAAD by NMNAT and
subsequently to NAD by NADS, as described for the NA to NAD pathway
earlier. QPRT represents the second rate-limiting step in the tryptophan
to NAD de novo synthesis pathway. It is expressed primarily in liver, kidney,
and brain (Magni et al., 2004; Rongvaux, Andris, Van Gool, & Leo,
2003). Deficiencies of QPRT activity are associated with neurological
problems, e.g., epilepsy and Huntington’s disease, likely reflecting the
neurotoxic properties of quinolinic acid (Feldblum et al., 1988; Foster,
Whetsell, Bird, & Schwarcz, 1985).
Niacin 97

3.2 NAD Recycling


Due to the continuous activity of NAD+-consuming enzymes, e.g.,
poly(ADP-ribose)polymerases (PARPs), sirtuins, and CD38, there is an
ongoing loss of NAD+ which is enzymatically cleaved usually with produc-
tion of nicotinamide as a reaction by-product. Simple dietary niacin intake
could not sustainably compensate for this NAD+ loss. Instead, nicotinamide
has to be efficiently recycled to NAD+. This is achieved by the consecutive
enzymatic activity of nicotinamide phosphoribosyltransferase (NAMPT)
and nicotinamide mononucleotide adenylyltransferase (NMNAT1–3), as
illustrated in Fig. 2 (green arrows, salvage pathway). The importance of this
pathway becomes clear when considering that levels of NAMPT in cells had
a larger effect on achievable NAD+ levels than concentrations of nicotin-
amide, as was demonstrated in cell culture models in which NAD+ levels
correlated with modulated expression of NAMPT (Yang et al., 2007).

3.3 NADPH Synthesis


NADH can be converted to NADPH phosphate (NADP) by the enzymatic
activity of NADH 20 -phosphotransferase or NADH kinase, which transfers a
phosphate group from ATP to NADH. NADPH is essential for many
reductive biosynthetic reactions in cells, e.g., fatty acid, steroid, and nucle-
otide synthesis, as outlined later.
During cellular metabolism, NADPH is regenerated from NADP via
several pathways, including the pentose phosphate pathway (aka hexose
monophosphate shunt), reductive reactions that reduce cytoplasmic oxalo-
acetate to pyruvate in the so-called pyruvate/malate cycle via the malic
enzyme, and NADP+-specific forms of isocitrate dehydrogenase (IDH)
and aldehyde dehydrogenase (ADLH) (Pollak, D€ olle, & Ziegler, 2007).

3.4 NAD and NADP Utilization


Enzymes utilizing NAD and NADP fall into several distinct categories (see
Fig. 3). The first category of enzymes utilize NAD[H] and the phosphory-
lated form NADP[H] as cofactors for cellular reduction and oxidation (redox)
reactions, in which NAD and NADP oscillate between the oxidized forms
(NAD+ and NADP+) and the reduced forms (NAD(H) and NADP(H)).
While the ratio of oxidized to reduced forms depends on the metabolic reac-
tion involved, the total amount of NAD or NADP is not affected.
The second group of NAD+-dependent enzymes represents a category
of enzymes that utilized only the oxidized form, i.e., NAD+, as a substrate in
reactions that require ADP-ribose moieties (see Fig. 3, red arrows). Such
98 James B. Kirkland and Mirella L. Meyer-Ficca

Purine
receptors
– Neurotransmitter

NADH
NADP+/NADPH
Redox reactions
- Cellular metabolism
Nicotinamide - Protection from
oxidative damage
adenosine dinucleotide
O
(NAD+)
O- NH2
+
O P O N
O

Recycling O
OH OH NH2
(NAMPT, N N
NMNAT) O P O N N
O- O

OH OH

Nicotinamide CD38,
“niacinamide” CD157
O MARTs Sirtuins
PARPs
NH2

N
– Inflammation – Longevity – Genomic stability
– Metabolic control – Longevity
(via calcium – Mitochondrial function – Epigenetics
signaling)
(via PTM control in (via DNA repair,
chromatin and chromatin and
transcriptional factors) transcription)
Fig. 3 NAD as essential cofactor for a wide variety of cellular processes. NAD is an essen-
tial enzymatic cofactor for most biochemical reactions in cellular metabolism processes
and is required to regenerate glutathione necessary to prevent and minimize cellular
damage due to naturally occurring reactive oxygen species (black circles, right upper cor-
ner). Additionally, NAD serves as a neurotransmitter, when it binds to and activates
purine receptors (black arrows, left upper corner). Three enzyme groups use NAD as
an enzymatic cosubstrate. The groups comprise MARTs, PARPs, and sirtuins (red arrows,
bottom part of diagram). Activity of these enzymes leads to a cleavage of NAD, which
causes loss of available NAD and the production of nicotinamide as a reaction
by-product. Increasing concentration of nicotinamide has inhibitory effects on PARPs
and sirtuins and reduces their activity (indicated by gray blocking arrows). Nicotinamide
is used as a substrate for NAD regeneration through activity of NAMPT and NMNAT in
the NAD salvage pathway (green arrow). Abbreviations: MART, mono(ADP-ribose)
transferase; NAMPT, nicotinamide phosphoribosyl transferase; NMNAT, nicotinamide
mononucleotide adenylyltransferase; PARP, poly(ADP-ribose)polymerase.
Niacin 99

reactions require enzymatic cleavage of NAD+ and therefore result in a


change of total NAD+ amounts that are available for all cellular reactions.
Enzyme families in this category comprise PARP enzymes (PARPs, also
known as ADP-ribose transferases (ARTDs)), sirtuins (members of the class
III NAD-dependent histone deacetylases, HDACs), NAD+-dependent
ADP cyclases CD38 and CD157, and indoleamine 2,3-dioxygenases
(IDOs).
The most recently discovered third function of NAD+ is as a ligand to a
small group of purine receptors, e.g., P2Y1 and P2Y11. Regulated NAD+
release occurs in neurosecretory cells, vascular and visceral smooth muscles,
urinary bladder, large intestine, as well as the brain (Breen, Smyth,
Yamboliev, & Mutafova-Yambolieva, 2006; Durnin et al., 2012;
Mutafova-Yambolieva, 2012; Mutafova-Yambolieva et al., 2007; Smyth,
Bobalova, Mendoza, Lew, & Mutafova-Yambolieva, 2004; Yamboliev,
Smyth, Durnin, Dai, & Mutafova-Yambolieva, 2009). NAD+-responsive
receptors are present on colonic cells, nucleated blood cells such as mono-
cytes, and endothelial cells of arteries, where NAD+ neurotransmitter-like
actions at purine receptors alter vascular tone.

3.4.1 Redox Reactions


NAD+ and NADP+ are indispensable cofactors in over 400 enzymatic redox
reactions central to most metabolic processes (https://www.cmescribe.com/
vitamin-dependent-gene-databases/) (Penberthy & Kirkland, 2012). The
following section provides a general and abbreviated overview highlighting
some of the most important biochemical processes that rely on sufficient
NAD and NADP availability, and hence indirectly on sufficient dietary nia-
cin consumption.
The redox couple NAD+/NADH [designated NAD(H) from here on]
primarily promotes oxidative and catabolic reaction, while NADP+/
NADPH [designated NADP(H)] most often drives reductive and anabolic
reactions. This difference is due to the wide difference in redox potentials
between those two redox couples. Under normal cellular conditions, most
of the molecules of the NAD(H) redox couple are present in the form of
oxidized NAD+, while most molecules in the NADP(H) redox couple
occur in the reduced form NADPH.
The oxidized NAD+ is necessary mostly for catabolism, e.g., in the
breakdown of glucose to pyruvate via glycolysis and the formation of acetyl
coenzyme A and subsequently complete catabolism in the Krebs cycle,
which generates NADH reduction equivalents that drive energy production
100 James B. Kirkland and Mirella L. Meyer-Ficca

via the respiratory chain in the mitochondria. In addition, NAD+ reduction


to NADH facilitates β-oxidation of fatty acids in the mitochondria. The
overview illustrates some select NAD+-dependent enzymes and their
specific enzymatic functions in diverse catabolic processes are shown in
Fig. 3. NAD+-dependent steps are central to energy production from
glucose via glycolysis, Krebs cycle, and anaerobic oxidation (steps 1–3 in
Fig. 4). Enzymes mediating central oxidative steps, such as glyceraldehyde
3-phosphate dehydrogenase, pyruvate dehydrogenase, IDH, and lactate
dehydrogenase, all require NAD+ as coenzyme. Various enzymes required
for β-oxidation of fatty acid processes during lipolysis also require NAD+ as a
cofactor, e.g., L-3-hydroxyacyl coenzyme A dehydrogenase and short-
chain hydroxyacyl coenzyme A dehydrogenase (step 4 in Fig. 4). Energy
generation via amino acid catabolism similarly is mediated by enzymes that
require NAD+ as a cofactor to catalyze oxidations steps, e.g., glutamate
dehydrogenase, branched chain a-ketoacid dehydrogenase, etc. (step 5 in
Fig. 4). This overview is far from being complete and is merely intended
to illustrate how central NAD+ is to most biochemical processes in the cells.
Both NAD+ and NADH/H+ are cofactors for some anabolic processes,
as illustrated in Fig. 5. If glucose levels are insufficient, cells can synthesis
glucose via gluconeogenesis. In this process, the enzymatic equilibrium of
three enzymes in the glycolysis pathway is reverted to allow them to utilize
the reductive energy of NADH/H+ to facilitate the synthesis of glucose
molecules starting from substrates such as lactate, pyruvate, and acetyl coen-
zyme A that have been generated by degradation of either amino acids or
fatty acids. Common enzymes with such a reversible enzymatic equilibrium
are lactate dehydrogenase, malate dehydrogenase, and glyceraldehyde
3-phosphate dehydrogenase (see step 1 in Fig. 5). NADH/H+ cofactors also
facilitate the synthesis of fat via the synthesis of glycerol 3-phosphate (step 2
in Fig. 5) and are required for the synthesis of dihydrotestosterone from
testosterone as a cofactor for the 5-α-reductase enzyme that mediates this
last conversion (step 3 in Fig. 5).
NADPH, the reduced form of NADP+, is an essential cofactor for many
anabolic biochemical processes, e.g., the synthesis of fatty acids and choles-
terol (see Fig. 6). It also required to support the reductive processes necessary
in cytochrome P450-mediated functionalization reactions, glutathione/fatty
acid hydroperoxidases, and thioredoxin defense against reactive oxygen
stress, and as a substrate for immune response-associated oxidative defense
reactions (Pollak et al., 2007).
During fatty acid synthesis, two molecules of NADPH are required for
each acetyl group that is added and reduced during lipogenesis. Enzymes that
Niacin 101

Fig. 4 Overview of biochemical redox reactions in catabolic processes that utilize NAD+
as a cofactor. This list shows important catabolic processes that lead to reduction of
NAD+ to NADH/H+. The general pathway name is indicated next to the numbers. Spe-
cific reactions occurring during the indicated general process are shown by name of
reaction intermediated (in black); enzymes that catabolize the reaction are indicated
in blue; relevant amino acid examples are indicated in green. Abbreviations: AA, amino
acid; CoA, coenzyme A.

are active in the process of fatty acid synthesis and require NADPH as a
coenzyme include β-ketoacyl ACP reductase and enoyl ACP reductase (step
1 in Fig. 6). In liver cells, the availability of NADPH therefore depends on
the activity of ongoing fatty acid synthesis processes.
102 James B. Kirkland and Mirella L. Meyer-Ficca

Fig. 5 Overview of biochemical redox reactions in anabolic processes that utilize NADH
or NAD+ as a cofactor. This list shows important anabolic processes that are associated
with either the reduction of NADH/H+ to NAD+ (steps 1–3) or the reduction of NAD+ to
NADH/H+ (step 4). The general pathway name is indicated next to the numbers. Specific
reactions occurring during the indicated general process are shown by name of reaction
intermediated (in black); enzymes that catabolize the reaction are indicated in blue.

Cholesterol synthesis is a complex biochemical synthesis process that


requires consumption of NAD+ as well as NADPH. Early in this process,
the enzyme 3-hydroxy-3-methylglutaryl CoA reductase requires NADPH
as a cofactor to synthesize mevalonate from 3-hydroxy-methylglutaryl CoA.
Mevalonate is one of the basic building blocks for cholesterol synthesis. Sev-
eral pharmacological substances that prevent this enzymatic step are admin-
istered to human patients as cholesterol-lowering drugs, e.g., lovastatin,
atorvastatin, and simvastatin. Interestingly, combinations of such drugs with
niacin are used to treat dyslipidemia. Later sections of this chapter will
review this in more detail. In a similar fashion, the enzyme squalene
synthase, which catalyzes the step from farnesyl pyrophosphate to squalene,
requires NADPH. It is inhibited by the drug squalestatin. Several biochem-
ical pathways can be used to synthesis cholesterol and all require NADPH as
a cofactor for the respective enzymes (see step 2 in Fig. 6).
Niacin 103

Fig. 6 Overview of biochemical redox reactions in anabolic processes that utilize


NADPH as a cofactor. This list shows some important anabolic processes that are asso-
ciated with either the reduction of NADPH/H+ to NADP+. The general pathway name is
indicated next to the numbers. Specific reactions occurring during the indicated general
process are shown by name of reaction intermediated (in black); enzymes that catab-
olize the reaction are indicated in blue.

Further important biochemical processes that require NAD+ and


NADPH as redox cofactors are various synthesis reactions that originate
from cholesterol: the synthesis of bile acids is initiated by conversion of
cholesterol to 7-α-hydroxycholesterol by 7-α-hydroxylase under the
104 James B. Kirkland and Mirella L. Meyer-Ficca

consumption of NADPH and O2 (step 3 in Fig. 6). Cholesterol also is the


basic building block for the generation of steroid hormones.
High levels of NADPH are present in red blood cell due to the fact that
red blood cells lack mitochondria, and therefore have to rely on the pentose
phosphate pathway for their energy metabolism. The resulting abundance of
NADPH supplies these cells with the important cofactor that allows ongoing
activity of glutathione reductase, which allows glutathione/fatty acid
hydroperoxidase enzymes to reduce oxidative damage within the iron
and oxygen-rich environment of red blood cells.
NAD(H) and NADP(H) also play a central role in the metabolism of eth-
anol and other alcohols. Ethanol is popular as a mood-altering compound
that is consumed mostly as wine, beer, and spirits. Small quantities of alcohol
intake have been associated with health benefits, while consumption of large
ethanol quantities causes severe physiological complications, i.e., liver cir-
rhosis, alcoholic fatty liver disease (AFLD), and hypoglycemia (Gaziano &
Manson, 1996). Through a series of consecutive enzymatic oxidation reac-
tions, ethanol is converted to acetaldehyde and finally to acetate. The initial
metabolic step, the conversion of ethanol to acetaldehyde, can be catalyzed
by either the enzyme alcohol dehydrogenase with production of NADH or
enzymes of the microsomal ethanol-oxidizing system, i.e., cytochrome
P450 complex enzymes, with consumption of O2 and NADPH/H+, and
leakage of superoxide. Higher ethanol doses, and chronic exposure, will
force more ethanol through the P450 route, leading to greater risk of hepatic
injury. Also, detoxification of the resulting acetaldehyde to acetate requires
the enzymatic activity of ADLH with NAD+ as the enzymatic cofactor.
Chronic ethanol consumption has been associated with a decrease in blood
NAD+ levels, and since the accumulation of acetaldehyde is associated with
the unpleasant consequences often experiences with alcohol consumption
“hangover,” supplemental niacin intake is thought to help with more effi-
cient ethanol metabolism. If ethanol consumption occurs to a degree at
which ethanol metabolism leads to markedly decreased NAD+ and increased
NADH/H+ levels, the enzymatic equilibrium in NAD-dependent enzymes
is shifted. Such a shift in NAD to NADH ratio results in a shift of the lactate
dehydrogenase product equilibrium from pyruvate to lactate, which results
in inhibited gluconeogenesis, and consequently can cause hypoglycemia.
A similar shift from α-ketoglutarate to isocitrate is observed for IDH, and
from oxaloacetate to malate for malate dehydrogenase. Together with a
direct NADH concentration-dependent inhibition of the enzyme
α-ketoglutarate dehydrogenase, this results in an inhibition of energy
Niacin 105

production via the Krebs cycle in the liver tissues that are detoxifying the
consumed ethanol.
Both NAD(H) and NADP(H) are essential to maintain the cellular respi-
ratory chain within all cells. Within cells, NAD and NADP are present at
higher concentrations in the mitochondria than the rest of the cell. This
compartmentalization explains why mitochondria-rich tissues, e.g., heart,
contain more NAD than tissues with fewer mitochondria, e.g., liver.
Absolute tissue content of NADP(H) and NAD(H) varies widely. Liver
contains about double the amounts of NAD(H) than NADP(H)
(800 nmol/g fresh rat liver tissue weight compared to 420 nmol/g tissue
weight), while skeletal muscle contains closer to 30 times more NAD(H)
than NADP(H) (590 nmol/g NAD(H) and 30 nmol/g NADP(H)).
Taken together, the role of NAD in redox metabolism relies heavily on
the ratio of oxides to reduced molecules, i.e., how many of the NAD(H)
molecules are present as NAD+ vs NADH/H+, since this ratio can cause
a significant shift in the product equilibrium of enzymatic reaction. This
can be seen, e.g., for lactate dehydrogenase, which produces mostly lactate,
if there is excessive NADH, while the presence of excessive oxidized NAD+
allows for the predominant formation of pyruvate. The ratio of NAD+ and
NADH/H+ is influenced (and influences) not only by the activity of a single
or few enzyme(s), but is the result of the sum of cellular metabolic processes
(glycolysis, citric acid cycle, fatty acid synthesis, mitochondrial respiratory
chain, etc.). The redox state of NAD(H) therefore reflects a complex read-
out of ongoing biochemical energy metabolism in a given cell type and sit-
uation, making NAD(H) an ideal candidate indicator molecule to monitor
cellular metabolism, due to its capability to integrate information regarding
nutritional quality, overall energy intake, and energy expenditure.

3.5 NAD as Substrate


Supporting the notion that NAD(H) can function as an indicator of the
cellular metabolic environment is the fact that NAD+ is the substrate for
enzymatic reactions that permit modulation of cellular epigenetic informa-
tion. NAD+, the oxidized form of NAD(H), is the form whose cellular
concentration varies most significantly and whose concentration most
closely reflects metabolic situations such as insufficient dietary niacin intake,
malnutrition, or excessive exercise. NAD+ levels fluctuate not only with
nutritional intake and exercise but also with circadian cues (Nakahata
et al., 2009). The close relationship between NAD+ availability and energy
106 James B. Kirkland and Mirella L. Meyer-Ficca

metabolism has resulted in studies testing if increased muscular NAD+ syn-


thesis could promote oxidative phosphorylation, and prompted thoughts of
modulating the NAD+ pool for therapeutic purposes, e.g., using NAD+
supplementation to improve energy metabolism in sporting performance
(Cantó, Menzies, & Auwerx, 2015; Frederick et al., 2015).
At the same time, activity of the NAD+-consuming enzymes will
directly control NAD+ concentrations and therefore directly change meta-
bolic processes in response to external exposures and situation that activate
these NAD-consuming enzymes. Several families of enzymes continuously
consume NAD+. As mentioned earlier, the NAD+-consuming enzyme
families comprise sirtuins, PARPs, ADP cyclases, and IDO. The degree
of NAD+ consumption is dependent on several factors, e.g., the presence
of external DNA-damaging agents. The enzymatic activity of the major
PARP enzyme, the nuclear enzyme PARP1, is strongly activated in
enzymes that bind to DNA strand breaks. Activation of PARP1 induces
enzymatic cleavage of NAD and utilization of the ADP-ribose moiety to
form poly(ADP-ribose) (PAR). The degree of enzymatic activation corre-
sponds to the degree of DNA damage. Moderate levels of PAR formation
facilitate efficient DNA repair and prevent cell cycle progression to prolif-
eration of abnormal cells that could lead to carcinogenesis (Masutani &
Fujimori, 2013; Masutani, Nakagama, & Sugimura, 2003). Exposure to
severe DNA damage will lead to PARP1 hyperactivation. This results in
dramatically increased cellular NAD+ consumption, and to a lethal degree
of cellular NAD+ depletion. Taken together, cellular NAD+ concentration
varies directly as a consequence of exposure to any kind of external DNA-
damaging conditions, e.g., radiation or exposure to alkylating agents
(Burkle & Virag, 2013; Fouquerel & Sobol, 2014; Kupper, van Gool, &
Burkle, 1995; Masutani et al., 2000; Meyer, Muller, Beneke, Kupper, &
Burkle, 2000; Schreiber et al., 1995; Van Gool et al., 1997). Reduced
NAD+ availability will not only change cellular metabolic processes but
can also result in altered gene expression, DNA repair, apoptosis, cell death,
or carcinogenesis. This example illustrates the close interdependence of cel-
lular levels of NAD+ not only on metabolic processes and the nutritional
situation but also on additional environmental conditions, e.g., to any kind
of exposure, that modulate the activities of individual NAD+-consuming
enzymes.

3.5.1 ADP-Ribosyl Cyclases


The NADase/ADP-ribosyl cyclase enzyme CD38 and its homolog CD157
use NAD+ and NADP+ to generate cyclic ADP-ribose (Aarhus, Graeff,
Niacin 107

Dickey, Walseth, & Lee, 1995; Ferrero, Lo Buono, Horenstein, Funaro, &
Malavasi, 2014; Guse & Lee, 2008). Both enzymes are highly expressed in
neutrophil cells and endothelial cells where they participate in immune
response-associated signal transduction. Thus, they were long regarded as
an immune cell “activation marker.” This view has evolved significantly
in the last years to CD38 as a multifunctional molecule (Quarona et al.,
2013). The synthesized cyclic ADP-ribose functions as a signaling molecule
that triggers an intracellular signaling cascade and results in Ca2+ elevation,
likely by activating receptors in the endoplasmic/sarcoplasmic reticulum
(Pollak et al., 2007). Immunocytes deficient in CD38 display a defect in
migration to inflammatory sites and render mice with such a defect more
susceptible to infections (Partida-Sánchez et al., 2001; Partida-Sanchez
et al., 2007; Schuber & Lund, 2004). Interestingly, CD38-deficient mice
retain higher NAD+ levels and are protected against obesity and metabolic
syndrome (Barbosa et al., 2007), potentially through changes in NAD+
levels, which then alter other enzymes, like the sirtuins (Escande et al.,
2013). CD38 was recently identified as a major factor contributing to the
so far enigmatic observation that NAD+ levels steadily decline with age
in a nutrition-independent fashion, and that this decline causes mitochon-
drial dysfunction (Camacho-Pereira et al., 2016; Schultz & Sinclair, 2016).
CD38/157-induced cyclic ADP-ribose and calcium release are also
involved in controlling oxytocin secretion and thus play a regulatory role
in social behaviors. In addition to support from animal models, a human sur-
vey has shown that expression levels and polymorphisms in these genes have
been associated with the Autism Quotient of individuals. These observations
provide some mechanistic insight to the appearance of dementia in pella-
grins, and the rapid disappearance of psychoses following niacin supplemen-
tation (Chong et al., 2017).

3.5.2 Poly(ADP-Ribose) Polymerases (PARPs) or ADP-Ribose


Transferases (ATRDs)
PAR formation, using NAD+ as a substrate, has been known for more
than 50 years. Chambon and colleagues first observed NAD+ depletion
and PAR formation upon DNA damage (Chambon, Weill, & Mandel,
1963). Most of the initial studies focused on the most prevalent and most
active PAR-forming enzyme, PARP1, whose central role in DNA repair
and the maintenance of genomic stability shaped the general and prevailing
assessment of poly(ADP-ribosyl)ation as the “guardian angel of the genome”
(Chatterjee, Berger, & Berger, 1999; Jeggo, 1998). Subsequent studies rev-
ealed the existence of additional proteins with structural homology to PARP
108 James B. Kirkland and Mirella L. Meyer-Ficca

(Johansson, 1999), and eventually of a large protein family of related


enzymes (Ame, Spenlehauer, & de Murcia, 2004; Meyer, Meyer-Ficca,
Jacobson, & Jacobson, 2004; Meyer-Ficca, Meyer, & Jacobson, 2005).
So far, 17 enzymes have been identified as members of the human PARP
superfamily, and our knowledge on the evolutionary origins, protein
domains, and cellular functions of these proteins grows continually
(Aravind, Zhang, de Souza, Anand, & Iyer, 2015; Gupte, Liu, & Kraus,
2017; Kraus, 2015; Perina et al., 2014; Poltronieri & Miwa, 2016).
In general, members of the PARP superfamily generate ADP-ribose
posttranslational modification at specific sites in target proteins. Among
the members of the PARP family, only some enzymes turned out to be true
PARPs. This means that they cleave the glycosylic bond between nicotin-
amide and ribose in NAD+, release nicotinamide, attach the liberated
ADP-ribose unit to a given target protein, and in subsequent steps
keep adding more ADP-ribose units to the initially attached ADP-ribose,
thereby generating polymers of (ADP-ribose) (PAR) as posttranslational
modifications. Fig. 7A illustrates the chemical structure of PAR molecules
(Meyer-Ficca et al., 2005). Detailed structural analyses demonstrated
that PAR can comprise hundreds of ADP-ribose units, either as linear or
as branched molecules (Alvarez-Gonzalez & Jacobson, 1987; Juarez-
Salinas, Levi, Jacobson, & Jacobson, 1982; Kiehlbauch, Aboul-Ela,
Jacobson, Ringer, & Jacobson, 1993; Miwa et al., 1981; Miwa, Saikawa,
Yamaizumi, Nishimura, & Sugimura, 1979; Miwa & Sugimura, 1984).
In the majority of PARP family members, the enzymatic activity is lim-
ited to the initial first step, i.e., cleaving NAD+, releasing nicotinamide, and
attaching one single ADP-ribose unit to the given target protein. Since the
subsequent addition of several units does not occur and only mono(ADP-
ribose) (MAR) is attached to the target protein, these enzymes have to be
classified as functional mono(ADP-ribose)transferases (MARTs). The pres-
ence of an HYE amino acid motif in the catalytic center was identified as
required, but not sufficient for PAR formation, as it is present in all four
enzymes with true PARP activity, i.e., PARP1, PARP2, PARP5a, and
PARP5b (better known as tankyrase 1 or TNKS1 and tankyrase 2 or
TNKS2). Several in vitro and in vivo studies ascertained that other members
of the PARP family either exhibited mono(ADP-ribosyl)transferase activity
(PARP 3, 4, 6–12, 14–16) or no activity could yet be demonstrated exper-
imentally (PARP13) (Karlberg et al., 2015; Meyer-Ficca et al., 2015; Vyas
et al., 2014; Yang et al., 2017). The discrepancies between common names
and enzymatic properties led to an effort aiming to change the nomenclature
from PARPs to “ADP-ribosyl transferases” in order to prevent the obvious
Niacin 109

A B
O
OH OH
Poly(ADP-ribose) + DNA damage –DNA damage
CH2 CH2 C O (PAR)
O
O (ADP-ribose)n
O P O
Target protein NH2
O
(e.g. PARP1 N N
O P O
itself)
O O N N

HO O OH OH
O
PAR
O
O P O
NH2
O
O P O N N
PARG O O N N
(ADP-ribose)n–x H2O +
HO O OH
+ ADP-ribose O

HO HO O O
Nuclei
O P O
O O NH2
O
O P O N N
NH2 O P O
O
O O N N
N N
O P O
N N O HO OH
O
HO OH

C Nicotinamide O
O O

NH2
Salvage pathway, PAR as PTM
+
NAD regeneration N

NH2
PARP-1
Enzymatic (inactive) N
N
O
activity
NH2 N
+ O O N

NH2
NAD N O P O P O
N
O
+ OH OH
O O N
N N OH O OH
O P O P O OH
O
OH OH
PARP-1 1⬙ 2⬘
OH O OH
OH OH (activated) PARG

Free PAR

PARP-1
DNA strand
breaks Free ADPR NH2

N
N

O O N
N
O P O P O
O
OH OH
OH O
OH OH
OH OH

Fig. 7 Poly(ADP-ribosyl)ation by PARP1. (A) The chemical structure of a poly(ADP-


ribose) molecule with a single branch point is shown. Glycosylic bonds that can be
cleaved by poly(ADP-ribose)glycohydrolase (PARG, indicated in red) are indicated by
red arrows. (B) HeLa cells treated with 500 μM MNNG as DNA-damaging agent
accumulation amounts of PAR that can easily be stained with PAR-specific
antibodies (visible as red fluorescence signals in the left top panel), while PAR
levels in untreated cells cannot be visualized clearly (right top panel). Bottom panel:
Blue, DAPI-stained nuclei for the cells seen in the top panels. (C) Poly(ADP-ribose) is
rapidly synthesized upon PARP1 activation. PAR automodification of PARP1
temporarily inactivates the enzyme until PAR is degraded by PARG activity and PARP1
is liberated in the not automodified form. Abbreviations: DAPI, 2-(4-amidinophenyl)-1H-
indole-6-carboxamidine; MNNG, N-methyl-N0 -nitro-N-nitrosoguanidine.

discrepancy between the term “PARP” as the common name for most
members of the PARP superfamily that are limited to MART activity
(Hottiger, Hassa, Luscher, Schuler, & Koch-Nolte, 2010). A low general
acceptance of this novel nomenclature has led to the simultaneous
110 James B. Kirkland and Mirella L. Meyer-Ficca

occurrence of several names for the same enzyme in the current scientific
literature.
In PARP1, the most active and best-studied PARP, enzymatic activity
is rapidly activated upon detection of and binding to DNA strand breaks.
Activated PARP1 uses NAD to form large and branched PAR molecules,
attached to various target proteins, including PARP1 itself, where it tempo-
rarily inactivates enzymatic activity (see Fig. 7C). The effects of PAR are
normally short-lived due to rapid degradation by poly(ADP-ribose)
glycohydrolase (PARG). Treatment of culture cells with DNA-damaging
agents leads to a rapid increase in PAR that can be detected using
specific antibodies, e.g., in Fig. 7B, where nuclear PAR is visible as red
signals in the top-left panel, while untreated control cells have no visible
signals.
The degree of enzymatic activity of PARP1 after activation by DNA
breaks directly reflects the degree of DNA damage that is present. The
synthesized PAR temporarily blocks cell cycle progression in order to allow
sufficient DNA repair and to prevent proliferation of damaged or mutated
cells. Extensive DNA damage leads to hyperactivation of PARP1, which
causes severe NAD+ depletion and results in cell death. This process has long
been regarded as a cancer prevention mechanism that allows for elimination
of severely damaged, potentially malignant cells. In line with this concept are
studies that show that Parp1-deficient mice are more sensitive to radiation
exposure and display increased genomic instability upon gamma radiation
and exposure to DNA-damaging drugs and have an increased predisposition
to develop tumors (Trucco et al., 1999; Wang et al., 1995). At the same
time, preventing PARP1 hyperactivation with its associated NAD+ loss
and spike in PAR formation either in genetic knock out model, or pharma-
cologically with PARP inhibitors protects from cell death after inflamma-
tory processes or ischemia–reperfusion injuries [reviewed by Herceg and
Wang (2001) and Shall and de Murcia (2000)]. These varied functions of
PARylation have led to ongoing interest in the development and clinical
use of PARP inhibitors for cancer treatment, as well as spiked interest in
exploring the benefit of PARP inhibition in nononcological diseases
(Berger et al., 2017; Lord & Ashworth, 2017; Rajawat, Shukla, &
Mishra, 2017).
Research over the last years has broadly expanded our understanding of
cellular functions of ADP-ribosylation beyond these accepted roles in DNA
repair.
Niacin 111

It has grown from primarily being regarded as mechanism safeguarding


genetic information via facilitation of DNA repair to the emerging view of
ADP-ribosylation as a versatile posttranslational modification that is associ-
ated with many key biological processes. In addition to its undisputed major
role in DNA repair, ADP-ribosylation facilitates and regulates DNA repli-
cation, cell division, transcription and signal transduction, cellular responses
to stress, infections, and aging (Palazzo, Mikoč, & Ahel, 2017). This is
achieved through modulation of a wide array of protein/protein interactions
and through control of transcription and epigenetic information encoded in
the specific DNA/protein interactions in chromatin.
Table 3 briefly illustrates major functions identified to date for individual
enzymes in the PARP superfamily.

3.5.3 Sirtuins
One other group of enzymes that consume NAD+ by enzymatic cleavage
are sirtuins. These are members of the class III NAD-dependent HDACs.
Sirtuins moved to the limelight of scientific interest by the discovery of their
central role mediating life span extension in situations of caloric intake
restriction (Anderson, Bitterman, Wood, Medvedik, & Sinclair, 2003;
Cohen et al., 2004; Howitz et al., 2003). To date, they have remained at
the center of ongoing scientific interest as pharmacological targets with a
potential to slow down or ameliorate the aging process (Grabowska,
Sikora, & Bielak-Zmijewska, 2017; Pan & Finkel, 2017).
Mammals possess seven ubiquitously expressed sirtuin homologs,
Sirt1–7. The common feature of sirtuins as class III HDACs is that they con-
sume NAD+ to be catalytically active (see Fig. 8). Consequently, changes of
NAD+ availability as well as the NAD+/NADH ratio directly regulate their
activity, which makes them excellent sensors for both cellular energy status
and redox status (Michan & Sinclair, 2007).
As HDACs, one of their major sirtuin function is the enzymatic
deacetylation of lysine residues on target proteins, and SIRT1–SIRT3,
SIRT5, and SIRT6 indeed preferentially act as deacetylases (see Fig. 8,
top section). The reaction is a two-step process that consumes NAD+:
sirtuins cleave the nicotinamide moiety of an NAD+ molecule, which
releases the ADP-ribose moiety. In a subsequent step, the acetyl or acyl
group of the target protein is then transferred from directly to this
ADP-ribose, resulting in a deacetylated target protein and the formation
of 200 -O-acetyl-ADP-ribose (Tanner, Landry, Sternglanz, & Denu, 2000).
112 James B. Kirkland and Mirella L. Meyer-Ficca

Table 3 The PARP Enzyme Superfamily


Enzymatic
PARP Synonyms Product Currently Known Predominant Function
DNA repair (multiple pathways),
transcriptional regulation, chromatin
1 ARTD1 PAR
structure regulation, cell death control,
inflammation
2 ARTD2 PAR DNA repair, transcriptional regulation,
chromatin structure regulation
3 ARTD3 MAR Transcriptional regulation
ARTD4, vault
4 MAR
PARP
ARTD5, Telomere length regulation, stress signal
5a PAR
tankyrase 1 transduction
ARTD6, Telomere length regulation, protein
5b PAR
tankyrase 2 degradation
6 ARTD17 MAR
ARTD14,
7 MAR TCDD-mediated toxicity, cell division
tiPARP
8 ARTD16 MAR Nuclear membrane integrity
9 ARTD9, BAL1 MAR Cell migration, cytoskeleton
10 ARTD10 MAR Signal transduction
11 ARTD11 MAR Nuclear membrane integrity
ARTD12,
12 MAR Cytoplasmic stress response
ZC3HDC1
ARTD13,
13 ZC3HAV1, Unknown miRNA-mediated gene silencing
ZAP
Cell migration, cytoskeleton, signal
14 ARTD8, BAL2 MAR
Transduction
15 ARTD7, BAL3 MAR Cytoplasmic stress response
16 ARTD15 MAR Unfolded protein response
Overview of current nomenclature, known enzymatic activity, and primary function.

The remaining sirtuins, SIRT4 and SIRT6, preferentially utilize NAD+ to


transfer ADP-ribose units to target proteins, an enzymatic property that
SIRT2 and SIRT3 execute in addition to their deacetylase properties. In
this reaction, the sirtuin enzyme cleaves NAD, releases the nicotinamide
Niacin 113

- Acetyllysine - Lysine
Acetylated N
Deacetylated
O
protein protein N
N
O O N

O
O
O P O P O
Deacetylation O O
(SIRT1, SIRT2, SIRT3, O O O O N
N
SIRT5, SIRT6) O O N
NH

N
H

O O 2’-O-Acetyl-ADP-ribose N
OH

O
+
H

O
O

NAD+ Salvage pathway


O

N H Nicotinamide
(NAMPT, MNMAT)
–O

O
P

H
O
O

N
P

O
O
H

N-H
N
HO

N
O
H

ADP-ribosylation
(SIRT2, SIRT3, SIRT4, SIRT6)
N
N
N
+ N
N
H H O
O
Protein Protein O
H
H O
O
H O
O O P
H O
O P O
O HH O
O

Fig. 8 Sirtuin-mediated activities. Sirtuins SIRT1, 2, 3, 5, and 6 utilize NAD+ to remove


posttranslational acetylation marks from target proteins, producing O-acetyl-ADP-
ribose and nicotinamide (top). Activities of SIRT4 and SIRT5 are limited to transfer of
ADP-ribose moieties to target proteins (bottom).

moiety, and enzymatically transfers the liberated ADP-ribose moiety to tar-


get proteins (see Fig. 8, bottom section). Functionally SIRT2, SIRT3,
SIRT4, and SIRT6 thus are MARTs.
As an overall consequence of collective sirtuin activity, epigenetic infor-
mation is changed, e.g., not only in the form of altered chromatin structure
following histone deacetylation but also in the form of direct changes of
transcription factor activities.
Among the sirtuin target proteins are transcription factors, e.g., FOXO1,
PGC-1α, and HIF-1α, i.e., factors that regulate oxidative metabolism, anti-
oxidant defense, and gene expression in the mitochondria both directly and
indirectly. Consequently, reduced sirtuin activity, e.g., as expected in
114 James B. Kirkland and Mirella L. Meyer-Ficca

situations of NAD+ paucity, will reduce the cell’s ability to perform oxida-
tive metabolism in the mitochondria (Imai & Guarente, 2016). The presence
of multiple sirtuin proteins, each with specific subcellular locations and tar-
get protein preferences, allow sirtuin activity to influence a variety of cellular
processes in the organisms, e.g., myogenesis, gluconeogenesis and insulin
secretion, adipogenesis, DNA repair, development, and senescence (Jiang,
Liu, Chen, Yan, & Zheng, 2017; Michan & Sinclair, 2007; van de Ven,
Santos, & Haigis, 2017; Vazquez, Thackray, & Serrano, 2017).
Sirtuins are functionally connected to NAD+ on several levels. Sirtuins
directly depend on NAD+ availability to be enzymatically active. Their
activity increases with rising NAD+ concentrations, while increasing levels
of NADH and nicotinamide directly inhibits sirtuin enzymatic activity. This
regulation allows sirtuins to directly sense and respond to cellular NAD+
concentrations and cellular redox status. At the same time, sirtuins compete
with other NAD+-consuming enzymes for available NAD+. PARP1 is
among the enzymes that have a high capacity for NAD+ utilization and
can deplete cellular NAD+ to levels that change the activity of sirtuins. Gen-
erally, tissue concentrations of NAD+ exceed the levels that sirtuins need for
activity. It still is conceivable that subcellular NAD+ concentrations in the
organelles vary and are in ranges that limit sirtuin activity; it is also possible
that DNA damage-induced activation of PARP1 depletes NAD+ to a
degree where it actively curbs sirtuin activity. On the other hand, sirtuin
activity directly downregulates PARP1 activity via deacetylation, pointing
to a complex situation where PARP1 and sirtuins compete for NAD+
resources.
A third link between sirtuins and NAD+ is provided by the reported con-
nection between sirtuins, the circadian transcription factors CLOCK/
BMAL1, and their regulator effect on expression of NAMPT, the enzyme
that is central for mammalian NAD+ synthesis in the NAD salvage pathway.
Together these factors form a circadian feedback loop that allows for the cir-
cadian oscillation of NAD+ directly connected to the activity of sirtuins
SIRT1 and SIRT6 (Asher et al., 2008; Belden & Dunlap, 2008;
Kritikou, 2008; Nakahata et al., 2008, 2009). An interesting extension of
the NAD+/circadian rhythm connection is the finding that liver PARP1
activity also oscillates in a daily manner and that is regulated by feeding
(Asher et al., 2010; Kumar & Takahashi, 2010; Peek et al., 2013;
Schuldt, 2010).
Collectively, there is a broad interplay of nutrition, metabolism, and cir-
cadian rhythm control, with NAD+ as a central player connecting dietary
Niacin 115

intake information with the individual facets of the transcription machinery


(Asher & Sassone-Corsi, 2015).
This close connection between NAD+ and sirtuin activity is seen as a
central component of the life span extending properties that sirtuins like, e.g.,
Sir2 (the yeast homolog to the human SIRT1) and SIRT1, exert. Mutations
that result in a loss of a functional NAD+ biosynthesis pathway cause signif-
icantly decreasing life span (Lin et al., 2002; Smith et al., 2000). A decrease in
NAD+ occurs also naturally during aging (Braidy et al., 2011; Massudi et al.,
2012), potentially exacerbating age-related health decline. Interestingly,
yeast with mutations in SIR2 orthologs that allow higher enzymatic activity
in an NAD-depleted environment was recently identified, raising the ques-
tion if polymorphisms in sirtuin genes that allow efficient enzymatic activity
in the presence of low NAD+ concentration have an evolutionary advantage
(Ondracek, Frappier, Ringel, Wolberger, & Guarente, 2017).
In conclusion, the collective activities of NAD+-dependent enzymes
(poly- and mono(ADP-ribosyl)transferases, ADP-ribosyl cyclases, and
sirtuins) provide mechanisms that integrate cellular signaling and with the
regulation of chromatin structure and epigenetic control, allowing for a spa-
tiotemporal fine-tuning of transcription and gene expression to adjust for
changed environmental conditions (see Fig. 9).

3.5.4 Indoleamine 2,3-Dioxygenase


Among the enzymes that potentially control NAD+ concentrations is IDO,
the enzyme mediating the initial and rate-limiting first step of the
kynurenine pathway. TDO and IDO activity controls the entry of the nor-
mally limiting essential amino acid tryptophan into the kynurenine pathway,
which results in decreased plasma tryptophan and increased NAD+. In con-
trast to mostly hepatic expression of TDO, IDO is expressed in various
tissues, e.g., neurological tissues, where it is subject to extensive regulation.
IDO activity can be activated by proinflammatory cytokines, e.g., inter-
feron gamma, IL-6, and nitric oxide. Immune responses like HIV/AIDS as
well as cancers and autoimmune disease can cause IDO overactivation and
consequently result in depletion of plasma tryptophan and increased concen-
trations of neurotoxic metabolites like quinolinic acid (Brown et al., 1991).
This was recently demonstrated in HIV/AIDS patient in low-income Afri-
can populations (Bipath, Levay, & Viljoen, 2015). Similarly, patients with
flaky dermatitis displayed metabolic changes, e.g., increased kynurenine/
tryptophan ratio and lowered tryptophan, which are consistent with
increased IDO activity (Maltos et al., 2015). In astroglia, metabolic, and
116 James B. Kirkland and Mirella L. Meyer-Ficca

Fig. 9 NAD availability regulates ADP-ribosylation and sirtuin-mediated epigenetic


changes. External factors, e.g., composition and quantity of food, metabolic parameters,
such as body condition, exercise, and caloric intake, as well as direct exposures to DNA
stressors (certain DNA-damaging toxicants, radiation, etc.) control the activity of the
sirtuins and ADP-ribose-transferase, two powerful epigenetic regulators. ADP-
ribosylation and sirtuin-dependent processes regulate the chromatin structure in a
large variety of ways, e.g., histone PTM modifications of histones and transcription fac-
tors, heterochromatin condensation, transcriptional regulation, and DNA repair and rep-
lication. Change in chromatin structure results in adjusted gene expression, which in
turn results in an overall adjustment of the physiological situation by controlling many
cellular processes, e.g., metabolism, cellular differentiation, and DNA damage response.
The extent of the enzymatic ADP-ribosylation and sirtuin response directly depends on
availably NAD+ levels (green arrows), allowing for a direct readout and integration of
complex nutritional and external factors into epigenetic gene expression control. At
the same time, the consumption of NAD+ in these processes changes NAD+ levels (dark
green arrows). The sum of the adjusted overall cellular responses, e.g., gene expression
response, will directly adjust cellular metabolism which in turn influences energy
metabolism and NAD+ oxidation status (light green arrow).
Niacin 117

neurotropic support cells, activation of IDO by the proinflammatory factor


IFN-gamma resulted in increased neurotoxic metabolites, but at the same
time maintained NAD levels via de novo synthesis by the kynurenine
pathway, which supports the hypothesis that one role of increased IDO
activity during inflammation is to maintain NAD+ levels (Grant &
Kapoor, 2003).
Local IDO upregulation causes cellular tryptophan starvation in certain
immune cells (e.g., T cells). This is required for the induction of immune
tolerance, as it is required to prevent fetal rejection during pregnancy, but
occurs similarly in autoimmune diseases and constitutes an immune-
suppressive mechanism that regulates the tumor microenvironment
(Muller & Prendergast, 2007; Munn & Mellor, 2016; Penberthy, 2007).
Induction of IDO and resulting immune suppression is also beneficial for
the suppression of graft-vs-host diseases and may have the therapeutic poten-
tial in a clinical setting (Jasperson et al., 2009). Inflammation induces IDO
activity and results in tryptophan depletion. Inflammation simultaneously
increases the production of iNOS-peroxynitrate activation of PARP activ-
ity, which results in chronic NAD+ depletion and persistent TNF-alpha
synthesis.
In addition to the connection to immune diseases, abnormal IDO reg-
ulation has been associated with cancer and neurological disorders and neu-
rodegenerative diseases, likely also via accumulation of neurotoxic
metabolites and depletion of neuroprotective tryptophan and NAD+
(Badawy, 2017). Increased tryptophan degradation via overactivation of
the kynurenine pathways has recently also been associated with cardiovas-
cular disease (Song, Ramprasath, Wang, & Zou, 2017).
In conclusion, the control of IDO activity by the immune system allows
for a local metabolic adjustment of innate and adaptive responses to inflam-
matory and immunological signals on a cellular and systemic level (Munn &
Mellor, 2013) and is associated with a control of NAD+ levels.

3.6 Turnover of PAR and MAR Modifications


Synthesis of PAR by PARPs, e.g., after exposure to DNA damage, leads to a
rapid and significant loss of cellular NAD+. Sirtuin activity, either as
deacetylation or as mono(ADP-ribosyl)ation (MAR), is a cause for
NAD+ loss in a similar manner. Both reactions release nicotinamide as a
reaction by-product, which will contribute to the NAD+ regeneration
via the NAD+ salvage pathway as described earlier (see Figs. 2, 3, 7, and 8).
118 James B. Kirkland and Mirella L. Meyer-Ficca

The additional moieties of NAD+, the ADP-ribose subunit that was


transferred to acceptor proteins during the enzymatic NAD+ cleavage reac-
tion, can undergo a similar regeneration. This regeneration is mediated by
enzymatic activities of PARG and ADP-ribosyl hydrolase-3 (ARH3), two
enzymes that cleave the linkage between ADP-ribose and the acceptor
proteins.

3.6.1 Poly(ADP-Ribose)Glycohydrolase
PARG releases PAR polymers and degrades the polymer into individual
ADP-ribose moieties by cleaving the O-glycosidic bond of the PAR chains.
Through alternative splicing, a single PARG gene gives rise to several
PARG protein isoforms of different sizes. Importantly, the different subcel-
lular localization allows PARG to degrade PAR molecules present in
different parts of the cell (Cortes et al., 2004; Koh et al., 2004; Meyer,
Meyer-Ficca, Whatcott, Jacobson, & Jacobson, 2007; Meyer-Ficca,
Meyer, Coyle, Jacobson, & Jacobson, 2004; Winstall et al., 1999). The
terminal ADP-ribose unit at the receptor glutamate amino acid residue in
a PAR-modified target protein can be removed by MACROD1 and MAC-
ROD2 proteins, and by TARG1/C6ord130 (Chen et al., 2011; Sharifi
et al., 2013). Deficiency of TARG1 is associated with neurological defects
comparable to the cytotoxic effects observed with neuronal PAR accumu-
lation (Andrabi et al., 2006; Sharifi et al., 2013; Yu et al., 2006). This empha-
sizes how important turnover and recycling of PAR and MAR signals are
for cellular functions.

3.6.2 ADP-Ribosyl Hydrolases (ARH1 and ARH3)


The ADP-ribosyl hydrolase family of enzymes consists of several
members. Of interest in this context are ARH1 and ARH3. ARH1 releases
the ADP-ribose subunit from proteins with mono-ADP-ribosylated argi-
nines, while ARH3 releases ADP-ribose moieties from PAR as well as from
O-acetyl-ADP-ribose, the by-product of sirtuin-mediated protein
deacetylation (Kasamatsu et al., 2011; Mashimo, Kato, & Moss, 2013, 2014).

4. NAD: RELEVANCE TO HEALTH AND DISEASE


Impaired niacin status can decrease total NAD(H), decrease NAD+,
shift the cellular NAD+ to NADH ratio, and change the ratios between
NAD(H) and NADP(H) pools. The ratios are likely important to health,
recalling that a shift toward reduced NADH occurs during aging.
Niacin 119

Furthermore, the shifted ratio also affects enzymatic reactions pertinent to


cellular metabolism. It changes the cells’ preference for oxidative phosphor-
ylation toward glycolysis, a process often observed in cancer cells. This
decreases pyruvate production and results in a net increased lactate forma-
tion. This increased glucose uptake and glucose to lactate fermentation is
known as the Warburg effect. The Warburg effect occurs even in the
presence of completely functional mitochondria and is known to promote
cellular growth and rapid proliferation, i.e., promoting tumorigenic prop-
erties (Liberti & Locasale, 2016).
Furthermore, the reduced NAD+ directly affects a large number of
NAD+-dependent signaling pathways discussed earlier (MARTs, PARPs,
sirtuins, ADP-ribose cyclases). Those pathways are involved in a large array
of biological processes, e.g., control of energy metabolism and mitochon-
drial functions, calcium homeostasis, control of exposure oxidative stress
(antioxidation mechanisms and generation of oxidative stress), control of
signaling and gene regulation pathways, immunological functions, aging,
and cell death (Ying, 2008). A common feature of the NAD+-dependent
pathways is the release of nicotinamide as a reaction by-product that can
be recycled to NAD (Opitz & Heiland, 2015).
In summary, disturbance of the NAD homeostasis and recycling
processes may potentiate an array of disorders that range from metabolic
syndrome, to cancer, neurological diseases, and age-associated decline.
The following section briefly illustrated some recent findings regarding
the connection between niacin, NAD, and clinical problems.

4.1 NAD Decline During Aging and Age-Related Disorders


Aging is a complex process caused by the interplay of many pathophysiolog-
ical processes. Two current theories on aging see either DNA damage
accumulation or mitochondrial deficiencies due to DNA damage following
nuclear to mitochondrial signaling as major contributors to aging. Both
processes are dependent on cellular NAD+ availability. Availability and
redox status of available NAD directly link DNA-damage repair with
mitochondrial function and nuclear to mitochondria signaling, indicating
the central role the adequate availability of niacin plays as a dietary precursor
of NAD for aging-related pathophysiology.
For many years, a connection has been proposed between NAD+ levels
and the link between aging and altered metabolism. This led to the term
“NAD world” as a relevant term in aging and with the view of NAD as
120 James B. Kirkland and Mirella L. Meyer-Ficca

a metabolic oscillator that regulates aging and metabolism (Imai, 2009a,


2009b, 2010, 2011; Imai & Guarente, 2014, 2016; Rehan, Laszki-
Szcza˛chor, Sobieszcza nska, & Polak-Jonkisz, 2014).
A significant and steady decline of NAD+ levels has been observed with
increasing aging in various species (Braidy et al., 2011; Camacho-Pereira
et al., 2016; Massudi et al., 2012; Schultz & Sinclair, 2016). In rats, a signif-
icant decline in tissue NAD and a shift in NAD+ to NADH ratio were
observed with aging, concomitantly with an increase in lipid peroxidation
and DNA damage as well as a decline in antioxidant capacity (Braidy et al.,
2011). A similar strong negative correlation between age and tissue NAD
content was observed in humans, with similarly increased lipid oxidation
and PARP activity (Massudi et al., 2012). Factors that contribute to this
age-associated NAD+ loss include increased activity of NAD+-consuming
enzymes, e.g., PARPs, CD38, and sirtuins (Jasper, 2013; Mouchiroud
et al., 2013; Schmeisser et al., 2013). Reducing NAD+ levels in worms
resulted in reduced life span, while pharmacological or genetic restoration
to normal NAD+ levels restored longevity and avoided age-associated met-
abolic decline (Mouchiroud et al., 2013).
Increased PARP activity in this context is thought to be a response to the
age-associated accumulation of DNA damage, and at the same time one of
the contributing factors responsible for the increased NAD+ catabolism seen
in aging. Furthermore, a recent study implicated an age-related increase in
the activity of the NAD-consuming enzyme CD38 with this age-associated
drop in NAD+ and mitochondrial dysfunction that was at least partially
dependent on a pathway regulated by SIRT3 (Camacho-Pereira et al.,
2016). This finding provided new clues to the long-known but puzzling
observation that NAD+ declines with age, independent of nutritional niacin
intake (Schultz & Sinclair, 2016). Keeping NAD+ levels up via a knockout
of CD38 protected from NAD+ decline and increased SIRT3 activity. Con-
sequently, improved mitochondrial function and increased glucose toler-
ance prevented aging-associated defects (Chini, Tarragó, & Chini, 2016).
Reduced NAD+ availability, in turn, prevents optimal sirtuin activity,
which is central to preventing age-associated changes in metabolic processes
and mitochondrial dysfunction (van de Ven et al., 2017). A simplified
scheme (Fig. 10) illustrates the vicious cycle of aging-related NAD decline,
and the exacerbating effects on cellular senescence, DNA damage, and met-
abolic dysfunction.
In model organisms, supplementation of NAD+ precursors, e.g., nico-
tinamide ribose, could increase net NAD+ synthesis and increase life span
Niacin 121

Fig. 10 Vicious cycle of age-dependent NAD+ loss. Aging (as cellular senescence) is
associated with exposure to increasing DNA damage and oxidative stress. Cellular
defense mechanisms against such stressors, e.g., PARPs and sirtuins, utilize NAD+.
Increased NAD+ use decreases NAD+ availability, which in turn increases aging-related
dysfunctions and promotes cellular senescence.

even in the absence of caloric restriction (Belenky et al., 2007). There are
current efforts to translate the growing knowledge about the correlation
between aging and declining blood NAD+ into clinical efforts to augment
intracellular NAD to prevent aging-associated disorders and decline
(Srivastava, 2016).

4.2 DNA Damage Accumulation


Aging is associated with an increase in basal DNA damage and increased rate
of mutations and cell death. PARP1 enzyme activation as a first response to
external DNA damage can facilitate DNA repair and, in the case of DNA
damage overwhelming the cellular repair capacity, induce apoptotic cell
death. This process requires utilization and cleavage of significant amounts
of NAD+ and therefore is indirectly dependent on nicotinamide as the die-
tary precursor for NAD+ in humans (for a recent review, see Surjana,
Halliday, & Damian, 2010). A moderate degree of cell damage can be
managed by the PARP1 systems while consuming a tolerable portion of
the cellular NAD+. Exposure to severe genotoxic stress in contrast causes
irreversible depletion of the cellular NAD+ content and results in cell
death (Schraufstatter, Hinshaw, Hyslop, Spragg, & Cochrane, 1986).
Supplementing NA to cultured human mononuclear blood cells improved
their genomic integrity after X-irradiation, as measured by increased viabil-
ity and reduced micronuclei formation. Supplemented cells also maintained
122 James B. Kirkland and Mirella L. Meyer-Ficca

higher cellular NAD+ levels, produced more PAR, and maintained higher
SIRT1 activity (Weidele, Beneke, & B€ urkle, 2017; Weidele, Kunzmann,
Schmitz, Beneke, & B€ urkle, 2010). An increase of intrinsic maximal PARP
activity correlates with the maximum life span observed on a species level
(Grube & B€ urkle, 1992). This indicates that species’ higher capability to effi-
ciently recognize and repair DNA damage has a beneficial effect to prevent
damage by the increased ROS insults that occur with aging.

4.3 Cell Death/Cancer


Genotoxic stress causes a rapid and oftentimes dramatic depletion of NAD+
from the nucleus and the cytoplasm, which triggers or at least contributes to
cell death. An increased mitochondrial NAD+ level can rescue cells from cell
death mediated by NAD+ depletion (Yang et al., 2007). This indicates that
mitochondrial NAD+ is a major determinant of apoptosis and presents the
intriguing possibility that boosting mitochondrial NAD via adequate diet
can determine organ health and disease.
Many of these protective pathways are dysregulated in cancer cells,
which makes NAD+ pathways intriguing targets for cancer therapeutics.
NAD+ depletion reduces energy synthesized by citric acid cycle activity
and oxidative phosphorylation and instead promotes Warburg-like metab-
olism focused on glycolysis. Recently studies have explored the effect of
well-known cancer therapeutics in combination with pharmacological
depletion of NAD+ levels. In many cases, there is a synergistic effect on can-
cer cell cytotoxicity. The combinations predispose cancer cells to oxidative
damage by disrupting their antioxidant defense system, increase cell death by
preventing DNA repair, and shift cell signaling pathways (e.g., SIRT1 and
p53) to a cytotoxic route (Kennedy et al., 2016).

4.4 Mitochondrial Dysfunction


4.4.1 Direct Connection Between Niacin, NAD, and Mitochondrial
Function
The correlation between aging and NAD+ levels and functionality are of
interest for mitochondrial health. The ratio of available NAD+ to NADH
regulates the activity of enzymes central to glycolysis, TCA cycle, fatty
acid oxidation, and fatty acid synthesis, i.e., pathways essential to cellular
metabolism. Aging and insufficient dietary niacin intake cause reduced
NAD+ availability and a shift toward more NADH that in turn leads to
reduced oxidative phosphorylation and impaired mitochondrial respiration,
while benefitting glucose consumption and Warburg effect.
Niacin 123

The reduced availability of NAD+ prevents optimal activity of the mito-


chondrial sirtuins (Sirt3–5). Sirt3–5 dependency on NAD+ allows them to
act as metabolic sensors that sense the metabolic state of the cell and translate
this info into mitochondrial regulation. Activity of the mitochondrial
sirtuins has been implicated in protection against age-related neuropathies,
insulin resistance, and cardiopathologies (van de Ven et al., 2017).
PARPs, the other major group of NAD+-consuming enzymes, also have
been identified as modulators of mitochondrial activity (Bai, Nagy, Fodor,
Liaudet, & Pacher, 2015). PARP hyperactivation causes a rapid loss of
mitochondrial potential and decreases mitochondrial oxygen consumption,
in particular upon ischemia–reperfusion injuries (Klaidman et al., 2003;
Módis et al., 2012). It can result in uncoupling of the electron transport chain
as well as superoxide production (Fang & Bohr, 2017; Fang et al., 2014).
Those observations have led to the suggestion of using PARP inhibitors
therapeutically to prevent cell death in ischemia–reperfusion scenarios, such
as myocardial infarction and stroke.
Examples of mitochondrial-related disease associated include glaucoma.
A recent study found an age-related decrease in retinal NAD+ levels
that makes neurons vulnerable to disease-related insults. Increased NAD+
content via oral administration or through gene therapy driving local
NAD+ production prevented glaucomas in 93% of eyes in the tested rodent
model (Williams et al., 2017).

4.4.2 Indirect Connection Between Niacin, NAD, and Mitochondrial


Function
Repletion of NAD+ can increase life span and—most importantly—health
span in animal models with defects in DNA repair response associated
with genetic aging disorders (such as, ATM/ataxia telangiectasia mutated,
XPA/xeroderma pigmentosum group A, CS/Cockayne syndrome). This
is thought to occur via the restoration of a defective mitophagy, the selective
clearance of defective mitochondria, as well as possible activation of general
autophagic proteins and through improved DNA repair (Croteau, Fang,
Nilsen, & Bohr, 2017; Fang & Bohr, 2017). One of the cellular problems
caused by these genetic defects is a loss of intracellular NAD+ that results
from persistent PARP1 activation due to increased DNA damage in the
mitochondria. This decreased NAD+ availability in turn causes other
NAD+-dependent enzymes, e.g., sirtuins, to suffer a loss of activity
(Croteau et al., 2017). SIRT1 is a multifunctional protein deacetylase whose
proper activity is central to longevity, metabolism, cellular senescence,
124 James B. Kirkland and Mirella L. Meyer-Ficca

genome maintenance, DNA repair, and inflammation. Hampering SIRT1


activity by restricting or making the rate limiting coenzyme NAD+
unavailable therefore leads to defects in mitochondrial homeostasis, to
increased production of reactive oxygen species, decreased DNA repair
capability and results in reduced cell survival.

4.5 Neurological Disorders


4.5.1 Age-Related Neurodegeneration
Although pellagrins with dementia display remarkable and rapid improve-
ment with niacin supplementation, suggesting a defect in cell signaling,
some brain pathologies are also observed with chronic pellagra. In the
aging brain, CNS decline underlies the progressive loss of cognitive, social,
and physical abilities that occur with aging. In this context, declining
sirtuin activity likely contributes to the aging process, opening up potential
opportunities for treating or preventing these age-associated dysfunctions
(Satoh, Imai, & Guarente, 2017).
Memory formation and circadian rhythms are closely intertwined, and
disruption of circadian rhythms, as it occurs, e.g., during aging, seems to
contribute to the age-associated memory loss. The close connection
between epigenetic changes, e.g., DNA methylation and sirtuin 1 activity,
and both memory formation and circadian rhythm, has given rise to the
concept that epigenetic changes in these processes contribute to the
age-associated cognitive decline (Deibel, Zelinski, Keeley, Kovalchuk, &
McDonald, 2015; Mazucanti et al., 2015).
Neurodegeneration in aging is often associated with mitochondrial
dysfunction. Recent studies have been able to link such mitochondrial-
associated neurodegeneration in animal models of several progeriatric
genetic diseases, e.g., CS and xeroderma pigmentosum group A (XPA) to
a decreased availability of mitochondrial NAD+ (Fang et al., 2014;
Scheibye-Knudsen et al., 2014).

4.5.2 Parkinson’s Disease


Parkinson’s disease (PD) is characterized by degeneration of neurons in
the substantia nigra of the brain stem, which leads to tremors associated with
insufficient dopamine synthesis. The complex etiology of the disease is not
completely understood, but includes genetic and environmental factors.
Dopamine is synthesized from tyrosine through several enzymatic steps that
among others requires NADH as a coenzyme.
Niacin 125

A direct link between dietary niacin intake, niacin index, and PD


progression was recently described by Wakade and Chong (2014) and
Wakade, Chong, Bradley, Thomas, and Morgan (2014). In PD patients,
low blood levels of NAD correlated with high expression levels of NA
receptor GPR109A PD patients compared to age-matched control individ-
uals. Dietary niacin supplementation normalized NAD levels and GPR109A
expression to levels seen in healthy control individuals and ameliorated
motor and cognitive functions in the PD patient (Wakade, Chong,
Bradley, & Morgan, 2015).
Recent genome-wide association studies have identified several poly-
morphic gene loci associated with altered risks or age of onset for PD
(Bandres-Ciga et al., 2016; Davis et al., 2016). Among the commonly found
genes is ACMSD, an enzyme central in the tryptophan to NAD+ metabolic
pathway that links PD risk to NAD metabolism, albeit though currently
unknown mechanistic details.
Familial forms of PD are associated with a gene mutation of PINK1 that
results in mitochondrial impairment, and subsequent neurodegeneration
due to mitochondrial deterioration. The mitochondrial disruption is associ-
ated with alteration of the NAD+ redox state and a reduction in NAD+
metabolites. Dietary supplementation with nicotinamide as well as genetic
suppression of PARP activity was able to prevent mitochondrial dysfunction
in an animal model of this pink1 mutation-associated form of PD and could
prevent neurodegeneration (Lehmann, Loh, & Martins, 2017).
Similar beneficial effects of exogenous niacin supplementation had
previously been observed in a human cellular PD model, in which niacin
supplementation protected from exogenously induced mitochondrial
defects and oxidative damage (Jia et al., 2008). In a transgenic Drosophila
PD model, Jia et al. (2008) found that niacin supplementation facilitated
improved motor function. Taken together, it appears that mitochondrial
dysfunction caused by NAD deficiency is becoming apparent as one of
the many factors contributing to the pathogenesis of PD, and that a potential
benefit of dietary niacin supplementation warrants further studies of this
treatment option.

4.5.3 Multiple Sclerosis


Multiple sclerosis (MS) is a disease characterized by progressive neu-
rodegeneration associated with inflammation. Current treatment options
target inflammation, but appear unable to prevent the neurodegeneration
126 James B. Kirkland and Mirella L. Meyer-Ficca

that underlies the progressiveness of the disease. Mitochondrial dysfunction,


associated with reduced NAD+ levels and potentially caused by
inflammation-associated oxidative stress, was found to contribute to the
neurodegeneration seen in MS (reviewed by Nimmagadda et al., 2017).
Neuroprotective effects preventing axon degeneration were observed in
neurons with increased activity of NMNAT, i.e., enzymes that increased
NAD+ and preserved ATP production and mitochondrial function in
those neurons (Araki, Sasaki, & Milbrandt, 2004; Park et al., 2013;
Yahata, Yuasa, & Araki, 2009), and in animal models overexpressing
SIRT1 (Nimmagadda et al., 2013). Both pathways point toward a central
role of NAD+ depletion in the pathogenesis of the MS-associated neu-
rodegeneration and indicate that pharmacological interventions designed
to increase the intracellular NAD+ content, e.g., via dietary niacin
supplementation, should be studied in the context of clinical applications.

4.5.4 Schizophrenia
While the pathophysiology underlying most schizophrenia cases is not well
understood, a recent study observed that chronically ill schizophrenia
patients had a significant reduction in NAD+/NADH ratio compared to
a matched healthy control group (Kim et al., 2017). The study provides evi-
dence that schizophrenia is linked to redox imbalance in the brain, pointing
toward a role of niacin intake and resulting NAD+ in schizophrenia etiology.
Additionally, full expression of pellagrous dementia is similar to schizophre-
nia, suggesting that some of the underlying changes in calcium signaling in
the niacin-deficient brain may be similar to schizophrenia, although schizo-
phrenia patients, on the whole, do not improve with niacin supplementation
(Prousky, Millman, & Kirkland, 2011).

4.5.5 Memory Loss


Ataxia telangiectasia (AT) in humans is associated with progressive neu-
rodegeneration. Treatments that replenished intracellular NAD+ in AT ani-
mal models (mouse and worms) improved neuromuscular function, delayed
memory loss, and extended the life span of the test animals. Mechanistically,
the intervention prevented DNA damage accumulation and mitochondrial
dysfunction, and at the same time stimulated neuronal DNA repair and
improved mitochondrial quality, pointing toward a therapeutic potential
of NAD-stabilizing treatment to prevent memory loss and neu-
rodegeneration (Fang et al., 2016).
Niacin 127

4.6 Skin Health and Skin Cancer Prevention


Severe niacin deficiency, as seen in clinical pellagra, is characterized by ery-
thematous rashes in sun-exposed areas, hyperkeratosis, and dermal fibrosis.
NAD+ restoration via niacin supplementation allows the skin to recover.
NAD deficiency in human keratinocytes causes reduced growth rates,
increases rates of apoptotic cell death, and leads to formation of reactive oxy-
gen species and increased rates of DNA damage (Benavente, Schnell, &
Jacobson, 2012). In addition, niacin restriction caused photosensitization
in skin cell lines, as illustrated by increased amounts of DNA damage, a
reduction in DNA damage-induced PAR formation, and alterations of
sirtuin expression levels (Benavente et al., 2012).
Skin cells are also prone to exhibit epidermal differentiation defects
following chronic UV skin exposure. This effect could largely be averted
by administration of pharmacological doses of a topically available NA deriv-
ative (Jacobson et al., 2007). A potential underlying mechanism for this
protective effect of niacin might be altered cell signal mediated by stimula-
tion of the NA-responsive G-protein-coupled receptors GPR109A and
GPR109B. Expression pattern analyses demonstrate the presence of func-
tional receptors in normal human primary and immortalized keratinocytes,
while squamous cell carcinoma (SCC) cells only have nearly nonfunctional
receptors available (Bermudez et al., 2011).
Using a mouse animal model, Gensler, Williams, Huang, and Jacobson
(1999) showed that nutritive administration of nicotinamide could
inhibit photocarcinogene. Conversely, mild niacin deficiency in a different
mouse study demonstrated an increased incidence of skin cancers after
UV-B treatment, indicating that already mild niacin deficiency similar to
what is frequently seen in humans might be sufficient to increase the skin
cancer risk. In clinical settings and human treatments, benefits of niacin
application for skin growth and differentiation have long been demonstra-
ted in the form of improved wound healing, treatment of psoriasis with
6-aminonicotinamide, smoothed skin surface structure, and reduced light
damage in skin cells (Gehring, 2004). In line with these findings, another
recent large-scale clinical trial on the protective role of oral nicotinamide
against skin cancer recurrence found that niacin intake was associated with
a reduced risk of SCC (Park et al., 2017). The same study demonstrated a
marginally positive association between niacin intake and the risk of basal
cell carcinoma and, in men only, with melanoma formation. Overall,
Park et al. (2017) concluded that niacin intake might be of potential benefit
128 James B. Kirkland and Mirella L. Meyer-Ficca

with respect to development of SCC, but likely not for prevention of BCC
or melanoma.

4.7 Cardiovascular Disorders and Niacin in Clinical Trials


In pharmacological doses, niacin intake can reduce myocardial infarction,
atherosclerosis, and stroke. The majority of these beneficial effects are
thought to be mediated by changes of lipoprotein composition. The ability
of pharmacological niacin intake to lower blood lipids, and to increase high-
density lipoprotein (HDL) cholesterol, i.e., the “good” cholesterol, has long
been recognized. Niacin was the first hypolipidemic drug that significantly
reduced cardiovascular clinical events and mortality in patients with cardio-
vascular disease. For current reviews on function and the use of niacin, see la
Paz et al. (2016) and Zeman et al. (2016).
Niacin in cultured human aortic cells increased levels of NADH and
reduced glutathione, which caused a reduction of reactive oxygen species
synthesis, low-density lipoprotein oxidation, and synthesis of TNF-alpha
and TNF-alpha-associated monocyte adhesion. This study thus indicated
that therapeutic niacin administration inhibits vascular inflammation, an
independent protective effect on the cardiovascular system that functions
in addition to niacin’s antiatherosclerotic effects on lipoprotein composition
(Ganji, Qin, Zhang, Kamanna, & Kashyap, 2009).
Early clinical studies consistently demonstrated that pharmacological
doses of niacin, either as monotherapy or in combination with bile acid
and cholesterol-sequestering drugs such as colestipol, significantly reduced
serum cholesterol and triglyceride levels and improved clinical outcomes:
3-year mortality rate in survivors of myocardial infarction was significantly
reduced, the frequency of coronary lesion was reduced, and atherosclerosis
regressed and coronary status improved (Blankenhorn et al., 1987; Brown
et al., 1990; Carlson, Danielson, Ekberg, Klintemar, & Rosenhamer,
1977; Carlson & Rosenhamer, 1988; Kane, 1990).
The Coronary Drug Project, a niacin monotherapy trial intended to
evaluate the frequency of recurrent myocardial infarction over a 6-year
interval and the associated 15-year total mortality, found significantly
reduced infarct and mortality rates in niacin monotherapy patients
(Canner, Furberg, & McGovern, 2006).
Increased levels of low-density lipoprotein cholesterol (LDL-C) and
lower-than-average levels of high-density lipoprotein cholesterol (HDL-
C) in patients with dyslipidemia were found to be independent risk factors
Niacin 129

for CVD (Barr, Russ, & Eder, 1951; Gordon, Castelli, Hjortland, Kannel, &
Dawber, 1977). This has led to the so-called HDL hypothesis that high
HDL protects from arteriosclerosis, and simultaneous correction of the
LDL-C to HDL-C ratio together with a general reduction of blood lipids
may provide reduced CV morbidity and mortality (reviewed, e.g., in
Vergeer, Holleboom, Kastelein, & Kuivenhoven, 2010). Niacin (nicotinic
acid) intake increases blood HDL-C levels and reduces LDL-C levels. While
the underlying mechanism for increased HDL-C is not yet completely
understood, increased production of apolipoprotein A-1 (ApoA-1) and
ATP-binding cassette subfamily A member 1 (ABCA1) contributes to the
effect (Lamon-Fava et al., 2008; Pang et al., 2014; Rubic, Trottmann, &
Lorenz, 2004; Wu & Zhao, 2009).
The unique property of niacin to correct the HDL-C:LDL-C ratio has
generated interest in the potential for combined niacin–statin therapies to
reduce blood lipid levels by statins and correct HDL-C:LDL-C ratios by
niacin in dyslipidemic patients. A study comparing safety and effectiveness
of combination formulations of extended release niacin with either a statin
or a bile acid sequestrant showed that both, niacin alone and its combination
preparations, effectively reduced LDL-C and triglycerides and increased
HDL-C (Guyton et al., 1998). The combination of niacin with statin
appeared to be most effective. Further trials (including HATS,
ARBITER-2, ARBITER-3, ARBITER-6, Oxford Niaspan Study,
AVANT, and COMPELL) corroborated that treatment with niacin in
combination with statins effectively reduced LDL-C levels and increased
HDL-C levels. In these trials, significantly reduced thickening of the carotid
intima media, a surrogate marker for atherosclerosis, and reduced stenosis
progression were also observed (Brown et al., 2001; deGoma et al., 2015;
Lee et al., 2009; McKenney et al., 2007; Taylor, 2004; Taylor, Lee, &
Sullenberger, 2006; Taylor et al., 2009). In a clinical trial called the “NIA
Plaque study,” patients with statin therapy-adjusted LDL-C blood levels
were supplemented with niacin or placebo. While the addition of niacin
to the statin therapy increased HDL-C levels, the reduction in internal
carotid artery wall volume was comparable in both treatment groups. This
indicates that it may be sufficient to control LDL-C levels in order to reduce
atherosclerosis independent of HDL-C levels (Sibley et al., 2013).
Further clinical trials addressed the growing concern that the predictive
value of surrogate measurements, e.g., arterial wall thickness, as indicators
for the risk of actual CVD events may be limited. This would mean that
reduction of blood lipids and prevention of atherosclerosis by supplementing
130 James B. Kirkland and Mirella L. Meyer-Ficca

niacin in a statin treatment regimen might not necessarily translate into


improved clinical outcomes (Hochholzer, Berg, & Giugliano, 2011).
Two trials (AIM HIGH and HPS2-THRIVE) therefore tested whether
supplementing a statin therapy with niacin is able to reduce the risk of
CV events in patients with atherogenic dyslipidemia below the levels
achieved by statins alone or not (The AIM-HIGH Investigators, 2011b,
2011c). Despite significantly reduced triglyceride and increased HDL-C
levels, the incidence of cardiovascular incidents (stroke, nonfatal myocardial
infarction, ischemic stroke, acute coronary syndrome, etc.) appeared not to
be reduced by adding niacin to the statin therapy during a 36-month follow-
up period (The AIM-HIGH Investigators, 2011a; The HPS2-THRIVE
Collaborative Group, 2014). Niacin augmentation was associated with
significantly increased rates of certain serious adverse events, leading to early
termination of one study (Anderson et al., 2014; Lloyd-Jones, 2014; The
HPS2-THRIVE Collaborative Group, 2014). Skin-related adverse effects,
diabetic complications and new onset diabetes, infections, gastrointestinal
and musculoskeletal problems, heart failure, and a 9% increased risk of death
were all observed. There is also a growing concern that there is no true causal
relationship between low HDL-C and cardiovascular risk, and that low
HDL-C levels should be considered a risk marker instead of a risk factor
(Voight et al., 2012). If this is the case, additional niacin treatment to correct
HDL-C levels cannot be expected to yield clinical benefit (Lloyd-
Jones, 2014).
Reports of other clinical trials utilizing such combination therapies
conclude that niacin could be safely added to statin or combined statin/
cholesterol sequestrator regimens, but recommend monitoring of blood
glucose levels as an indicator of intermittent occurrence of transient diabetes
in patients on niacin (Fazio et al., 2010; Guyton et al., 2012).
The increased incidence of diabetes in patients in the niacin-
supplemented groups raised the question if the benefit of niacin therapy
depended on metabolic status of patients. While the niacin-mediated cor-
rection of plasma lipid levels occurs regardless of the presence of metabolic
syndrome or diabetes mellitus (Ooi et al., 2015), evidence is accumulating
from primary studies and meta-analyses that niacin treatment can indeed
induce hepatic insulin resistance and increase the risk of developing diabetes
(Blond et al., 2014; Goldie et al., 2015).
Vasodilation resulting in skin flushing and itching is a well-known side
effect of niacin intake that often leads to increased dropout rates in niacin
treatment clinical trial groups and to reduced patient compliance.
Niacin 131

A clinical trial focusing on safety and tolerability of prolonged-release


niacin found that a daily dose of 2000 mg over a period of 15 weeks
did not lead to significant hepatotoxicity or muscular adverse effects, but
to high incidence of skin flushing (>40% of patients), and a related rate
of 10% of patient that withdrew from the study (NAUTILUS study,
Vogt, Kassner, Hostalek, Peiter, & Steinhagen-Thiessen, 2006; Vogt,
Kassner, Hostalek, Steinhagen-Thiessen, & NAUTILUS Study Group,
2006). Coadministration of a selective prostaglandin D2 receptor subtype
1 antagonist, laropiprant, was shown to reduce the niacin-induced vasodi-
lation (Lai et al., 2007), to safely attenuate skin flushing and to improve
patient compliance (Hussein & Nicholls, 2010). In addition, niacin/
laropiprant combination therapy improves lipid levels across a range of
patient types, irrespective of sex, race, age, statin use, and presence of
coronary heart disease, including type 2 diabetes mellitus patients
irrespective of glycemic control (Bays et al., 2015, 2012; Bays, Shah,
Dong, McCrary Sisk, & Maccubbin, 2011).
The current therapeutic role of niacin for dyslipidemia treatment remains
unclear. The perceived negative outcome of the clinical trials that failed to
show a potential for niacin to further improve on clinical benefits of statin
treatments led to a slight decrease in current niacin therapy for improved
cardiovascular health (Bittner et al., 2015). The patients in those trials were
already treated to optimal low LDL-C levels with statins, and niacin favor-
ably changed HDL-C and triglyceride levels. Niacin therefore could be used
as adjuvant therapy to reduce atherogenic lipoprotein burden in patients that
cannot reach their target HDL-C and LDL-C values, or that have contra-
indications for taking statins or bile-acid sequestrants (Boden, Sidhu, &
Toth, 2014; Lloyd-Jones, 2014). There is also growing evidence that the
low HDL-C might not be a causal risk factor for cardiovascular disease,
but rather a marker of high risk for cardiovascular disease (often associated
with metabolic syndrome and insulin resistance) (Voight et al., 2012), in
which case pharmacologically increasing HDL-C could not be expected
to yield any clinical benefit. Current reevaluations of the AIM-HIGH
and the HPS2-THRIVE clinical trial showed that the perceived negative
outcome might not be justified due to several methodological flaws in
the studies. The reevaluation indicates that some of the beneficial
antiatherosclerotic effects of niacin supplementation can be due to niacin
supplementation boosting antioxidative and antiinflammatory mechanisms
(Zeman et al., 2016). It should be emphasized that these studies are using
doses in the range of 1–3 g/day of nicotinic acid, and these side effects are
132 James B. Kirkland and Mirella L. Meyer-Ficca

unlikely in studies supplementing 100 mg/day of nicotinamide or nicotin-


amide riboside, for example.

5. SUMMARY
The central role dietary niacin plays for cellular NAD levels and, thus
for most metabolic processes, is directly reflected in the variety and large
number of clinical problems associated with a suboptimal niacin status
and emphasizes how important the maintenance of a good niacin status is
for overall health. Niacin has a long and successful history as a safe vitamin
supplement, which makes it a promising candidate for therapeutic interven-
tions to ameliorate or treat many of the clinical problems described earlier,
and furthermore might prove beneficial in the prevention of health decline
associated with aging.

REFERENCES
Aarhus, R., Graeff, R. M., Dickey, D. M., Walseth, T. F., & Lee, H. C. (1995). ADP-ribosyl
cyclase and CD38 catalyze the synthesis of a calcium-mobilizing metabolite from NADP.
The Journal of Biological Chemistry, 270(51), 30327–30333.
Alvarez-Gonzalez, R., & Jacobson, M. K. (1987). Characterization of polymers of adenosine
diphosphate ribose generated in vitro and in vivo. Biochemistry, 26(11), 3218–3224.
Ame, J. C., Spenlehauer, C., & de Murcia, G. (2004). The PARP superfamily. BioEssays:
News and Reviews in Molecular, Cellular and Developmental Biology, 26(8), 882–893.
https://doi.org/10.1002/bies.20085.
Anderson, R. M., Bitterman, K. J., Wood, J. G., Medvedik, O., & Sinclair, D. A. (2003).
Nicotinamide and PNC1 govern lifespan extension by calorie restriction in Saccharomyces
cerevisiae. Nature, 423(6936), 181–185. https://doi.org/10.1038/nature01578.
Anderson, T. J., Boden, W. E., Desvigne-Nickens, P., Fleg, J. L., Kashyap, M. L.,
McBride, R., et al. (2014). Safety profile of extended-release niacin in the aim-high trial.
New England Journal of Medicine, 371(3), 288–290. https://doi.org/10.1056/
NEJMc1311039.
Andrabi, S. A., Kim, N. S., Yu, S. W., Wang, H., Koh, D. W., Sasaki, M., et al. (2006). Poly
(ADP-ribose) (PAR) polymer is a death signal. Proceedings of the National Academy of
Sciences of the United States of America, 103(48), 18308–18313. https://doi.org/10.1073/
pnas.0606526103.
Araki, T., Sasaki, Y., & Milbrandt, J. (2004). Increased nuclear NAD biosynthesis and SIRT1
activation prevent axonal degeneration. Science (New York, NY), 305(5686), 1010–1013.
https://doi.org/10.1126/science.1098014.
Aravind, L., Zhang, D., de Souza, R. F., Anand, S., & Iyer, L. M. (2015). The natural history
of ADP-ribosyltransferases and the ADP-ribosylation system. Current Topics in Microbiol-
ogy and Immunology, 384, 3–32. https://doi.org/10.1007/82_2014_414.
Asher, G., Gatfield, D., Stratmann, M., Reinke, H., Dibner, C., Kreppel, F., et al. (2008).
SIRT1 regulates circadian clock gene expression through PER2 deacetylation. Cell,
134(2), 317–328. https://doi.org/10.1016/j.cell.2008.06.050.
Asher, G., Reinke, H., Altmeyer, M., Gutierrez-Arcelus, M., Hottiger, M. O., &
Schibler, U. (2010). Poly(ADP-ribose) polymerase 1 participates in the phase entrain-
ment of circadian clocks to feeding. Cell, 142(6), 943–953. https://doi.org/10.1016/
j.cell.2010.08.016.
Niacin 133

Asher, G., & Sassone-Corsi, P. (2015). Time for food: The intimate interplay between nutri-
tion, metabolism, and the circadian clock. Cell, 161(1), 84–92. https://doi.org/10.1016/
j.cell.2015.03.015.
Aykroyd, W. R. (1971). Conquest of deficiency diseases achievements and prospects. 3. Pel-
lagra. Indian Journal of Medical Sciences, 25(3), 203–207.
Badawy, A. A.-B. (2017). Kynurenine pathway of tryptophan metabolism: Regulatory
and functional aspects. International Journal of Tryptophan Research: IJTR, 10.
1178646917691938. https://doi.org/10.1177/1178646917691938.
Bai, P., Nagy, L., Fodor, T., Liaudet, L., & Pacher, P. (2015). Poly(ADP-ribose) polymerases
as modulators of mitochondrial activity. Trends in Endocrinology and Metabolism: TEM,
26(2), 75–83. https://doi.org/10.1016/j.tem.2014.11.003.
Baker, H., DeAngelis, B., Holland, B., Gittens-Williams, L., & Barrett, T. (2002). Vitamin
profile of 563 gravidas during trimesters of pregnancy. Journal of the American College of
Nutrition, 21(1), 33–37.
Ball, H. J., Jusof, F. F., Bakmiwewa, S. M., Hunt, N. H., & Yuasa, H. J. (2014). Tryptophan-
catabolizing enzymes—Party of three. Frontiers in Immunology, 5(485). https://doi.org/
10.3389/fimmu.2014.00485.
Bandres-Ciga, S., Price, T. R., Barrero, F. J., Escamilla-Sevilla, F., Pelegrina, J., Arepalli, S.,
et al. (2016). Genome-wide assessment of Parkinson’s disease in a Southern Spanish
population. Neurobiology of Aging, 45, 213.e3–213.e9. https://doi.org/10.1016/j.
neurobiolaging.2016.06.001.
Barbosa, M. T. P., Soares, S. M., Novak, C. M., Sinclair, D., Levine, J. A., Aksoy, P., et al.
(2007). The enzyme CD38 (a NAD glycohydrolase, EC 3.2.2.5) is necessary for the
development of diet-induced obesity. FASEB Journal: Official Publication of the Federation
of American Societies for Experimental Biology, 21(13), 3629–3639. https://doi.org/10.
1096/fj.07-8290com.
Barr, D. P., Russ, E. M., & Eder, H. A. (1951). Protein-lipid relationships in human plasma. II.
In atherosclerosis and related conditions. The American Journal of Medicine, 11(4), 480–493.
Barrett-Connor, E. (1967). The etiology of pellagra and its significance for modern medicine.
The American Journal of Medicine, 42(6), 859–867.
Bays, H. E., Brinton, E. A., Triscari, J., Chen, E., Maccubbin, D., MacLean, A. A., et al.
(2015). Extended-release niacin/laropiprant significantly improves lipid levels in type
2 diabetes mellitus irrespective of baseline glycemic control. Vascular Health and Risk
Management, 11, 165–172. https://doi.org/10.2147/VHRM.S70907.
Bays, H., Shah, A., Dong, Q., McCrary Sisk, C., & Maccubbin, D. (2011). Extended-release
niacin/laropiprant lipid-altering consistency across patient subgroups. International
Journal of Clinical Practice, 65(4), 436–445. https://doi.org/10.1111/j.1742-1241.2010.
02620.x.
Bays, H. E., Shah, A., Lin, J., Sisk, C. M., Dong, Q., & Maccubbin, D. (2012). Consistency
of extended-release niacin/laropiprant effects on Lp(a), ApoB, non-HDL-C, Apo A1,
and ApoB/ApoA1 ratio across patient subgroups. American Journal of Cardiovascular Drugs:
Drugs, Devices, and Other Interventions, 12(3), 197–206. https://doi.org/10.
2165/11631530-000000000-00000.
Belden, W. J., & Dunlap, J. C. (2008). SIRT1 is a circadian deacetylase for core clock com-
ponents. Cell, 134(2), 212–214. https://doi.org/10.1016/j.cell.2008.07.010.
Belenky, P., Racette, F. G., Bogan, K. L., McClure, J. M., Smith, J. S., & Brenner, C. (2007).
Nicotinamide riboside promotes Sir2 silencing and extends lifespan via Nrk and Urh1/
Pnp1/Meu1 pathways to NAD+. Cell, 129(3), 473–484. https://doi.org/10.1016/j.cell.
2007.03.024.
Benavente, C. A., Schnell, S. A., & Jacobson, E. L. (2012). Effects of niacin restriction on
sirtuin and PARP responses to photodamage in human skin. PLoS One, 7(7). e42276
https://doi.org/10.1371/journal.pone.0042276.
Beretich, G. R. (2005). Do high leucine/low tryptophan dieting foods (yogurt, gelatin) with
niacin supplementation cause neuropsychiatric symptoms (depression) but not
134 James B. Kirkland and Mirella L. Meyer-Ficca

dermatological symptoms of pellagra? Medical Hypotheses, 65(3), 628–629. https://doi.


org/10.1016/j.mehy.2005.04.002.
Berger, N. A., Besson, V. C., Boulares, A. H., B€ urkle, A., Chiarugi, A., Clark, R. S., et al.
(2017). Opportunities for the repurposing of PARP inhibitors for the therapy of non-
oncological diseases. British Journal of Pharmacology. https://doi.org/10.1111/bph.13748.
Bermudez, Y., Benavente, C. A., Meyer, R. G., Coyle, W. R., Jacobson, M. K., &
Jacobson, E. L. (2011). Nicotinic acid receptor abnormalities in human skin cancer:
Implications for a role in epidermal differentiation. PLoS One, 6(5). e20487 https://
doi.org/10.1371/journal.pone.0020487.
Bieganowski, P., & Brenner, C. (2004). Discoveries of nicotinamide riboside as a nutrient
and conserved NRK genes establish a Preiss-Handler independent route to NAD + in
fungi and humans. Cell, 117(4), 495–502.
Bipath, P., Levay, P. F., & Viljoen, M. (2015). The kynurenine pathway activities in a sub-
Saharan HIV/AIDS population. BMC Infectious Diseases, 15(346) https://doi.org/10.
1186/s12879-015-1087-5.
Bittner, V., Deng, L., Rosenson, R. S., Taylor, B., Glasser, S. P., Kent, S. T., et al. (2015).
Trends in the use of nonstatin lipid-lowering therapy among patients with coronary heart
disease: A retrospective cohort study in the medicare population 2007 to 2011. Journal of
the American College of Cardiology, 66(17), 1864–1872. https://doi.org/10.1016/j.jacc.
2015.08.042.
Blankenhorn, D. H., Nessim, S. A., Johnson, R. L., Sanmarco, M. E., Azen, S. P., & Cashin-
Hemphill, L. (1987). Beneficial effects of combined colestipol-niacin therapy on coro-
nary atherosclerosis and coronary venous bypass grafts. JAMA, 257(23), 3233–3240.
Blond, E., Rieusset, J., Alligier, M., Lambert-Porcheron, S., Bendridi, N., Gabert, L., et al.
(2014). Nicotinic acid effects on insulin sensitivity and hepatic lipid metabolism: An in
vivo to in vitro study. Hormone and Metabolic Research ¼ Hormon- Und
Stoffwechselforschung ¼ Hormones Et Metabolisme, 46(6), 390–396. https://doi.org/10.
1055/s-0034-1372600.
Boden, W. E., Sidhu, M. S., & Toth, P. P. (2014). The therapeutic role of niacin in dys-
lipidemia management. Journal of Cardiovascular Pharmacology and Therapeutics, 19(2),
141–158. https://doi.org/10.1177/1074248413514481.
Bouma, G., van Faassen, M., Kats-Ugurlu, G., de Vries, E. G. E., Kema, I. P., &
Walenkamp, A. M. E. (2016). Niacin (Vitamin B3) supplementation in patients with
serotonin-producing neuroendocrine tumor. Neuroendocrinology, 103(5), 489–494.
https://doi.org/10.1159/000440621.
Braidy, N., Guillemin, G. J., Mansour, H., Chan-Ling, T., Poljak, A., & Grant, R. (2011).
Age related changes in NAD + metabolism oxidative stress and Sirt1 activity in wistar
rats. PLoS One, 6(4). e19194 https://doi.org/10.1371/journal.pone.0019194.
Breen, L. T., Smyth, L. M., Yamboliev, I. A., & Mutafova-Yambolieva, V. N. (2006). Beta-
NAD is a novel nucleotide released on stimulation of nerve terminals in human urinary
bladder detrusor muscle. American Journal of Physiology Renal Physiology, 290(2),
F486–495. https://doi.org/10.1152/ajprenal.00314.2005.
oer, S., Cavanaugh, J. A., & Rasko, J. E. J. (2005). Neutral amino acid transport in epi-
Br€
thelial cells and its malfunction in Hartnup disorder. Biochemical Society Transactions,
33(Pt. 1), 233–236. https://doi.org/10.1042/BST0330233.
Brown, G., Albers, J. J., Fisher, L. D., Schaefer, S. M., Lin, J.-T., Kaplan, C., et al. (1990).
Regression of coronary artery disease as a result of intensive lipid-lowering therapy in
men with high levels of apolipoprotein B. New England Journal of Medicine, 323(19),
1289–1298. https://doi.org/10.1056/NEJM199011083231901.
Brown, R. R., Ozaki, Y., Datta, S. P., Borden, E. C., Sondel, P. M., & Malone, D. G.
(1991). Implications of interferon-induced tryptophan catabolism in cancer, auto-
immune diseases and AIDS. Advances in Experimental Medicine and Biology, 294, 425–435.
Niacin 135

Brown, B. G., Zhao, X. Q., Chait, A., Fisher, L. D., Cheung, M. C., Morse, J. S., et al.
(2001). Simvastatin and niacin, antioxidant vitamins, or the combination for the preven-
tion of coronary disease. The New England Journal of Medicine, 345(22), 1583–1592.
https://doi.org/10.1056/NEJMoa011090.
Burkle, A., & Virag, L. (2013). Poly(ADP-ribose): PARadigms and PARadoxes. Molecular
Aspects of Medicine, 34, 1046–1065. https://doi.org/10.1016/j.mam.2012.12.010.
Camacho-Pereira, J., Tarragó, M. G., Chini, C. C. S., Nin, V., Escande, C., Warner, G. M.,
et al. (2016). CD38 dictates age-related NAD decline and mitochondrial dysfunction
through an SIRT3-dependent mechanism. Cell Metabolism, 23(6), 1127–1139.
https://doi.org/10.1016/j.cmet.2016.05.006.
Canner, P. L., Furberg, C. D., & McGovern, M. E. (2006). Benefits of niacin in patients with
versus without the metabolic syndrome and healed myocardial infarction (from the Cor-
onary Drug Project). The American Journal of Cardiology, 97(4), 477–479. https://doi.org/
10.1016/j.amjcard.2005.08.070.
Cantó, C., Menzies, K. J., & Auwerx, J. (2015). NAD+ metabolism and the control of
energy homeostasis: A balancing act between mitochondria and the nucleus. Cell Metab-
olism, 22(1), 31–53. https://doi.org/10.1016/j.cmet.2015.05.023.
Carlson, L. A., Danielson, M., Ekberg, I., Klintemar, B., & Rosenhamer, G. (1977). Reduc-
tion of myocardial reinfarction by the combined treatment with clofibrate and nicotinic
acid. Atherosclerosis, 28(1), 81–86.
Carlson, L. A., & Rosenhamer, G. (1988). Reduction of mortality in the Stockholm
Ischaemic Heart Disease Secondary Prevention Study by combined treatment with clo-
fibrate and nicotinic acid. Acta Medica Scandinavica, 223(5), 405–418.
Chambon, P., Weill, J. D., & Mandel, P. (1963). Nicotinamide mononucleotide activation
of new DNA-dependent polyadenylic acid synthesizing nuclear enzyme. Biochemical and
Biophysical Research Communications, 11, 39–43.
Chatterjee, S., Berger, S. J., & Berger, N. A. (1999). Poly(ADP-ribose) polymerase:
A guardian of the genome that facilitates DNA repair by protecting against DNA recom-
bination. Molecular and Cellular Biochemistry, 193(1–2), 23–30.
Chen, D., Vollmar, M., Rossi, M. N., Phillips, C., Kraehenbuehl, R., Slade, D., et al. (2011).
Identification of macrodomain proteins as novel O-acetyl-ADP-ribose deacetylases. The
Journal of Biological Chemistry, 286(15), 13261–13271. https://doi.org/10.1074/jbc.
M110.206771.
Chini, C. C. S., Tarragó, M. G., & Chini, E. N. (2016). NAD and the aging process: Role in
life, death and everything in between. Molecular and Cellular Endocrinology, 455, 62–74.
https://doi.org/10.1016/j.mce.2016.11.003.
Chong, A., Malavasi, F., Israel, S., Khor, C. C., Yap, V. B., Monakhov, M., et al. (2017). ADP
ribosyl-cyclases (CD38/CD157), social skills and friendship. Psychoneuroendocrinology, 78,
185–192. https://doi.org/10.1016/j.psyneuen.2017.01.011.
Cohen, H. Y., Miller, C., Bitterman, K. J., Wall, N. R., Hekking, B., Kessler, B., et al.
(2004). Calorie restriction promotes mammalian cell survival by inducing the SIRT1
deacetylase. Science (New York, NY), 305(5682), 390–392. https://doi.org/10.1126/
science.1099196.
Cortes, U., Tong, W. M., Coyle, D. L., Meyer-Ficca, M. L., Meyer, R. G., Petrilli, V., et al.
(2004). Depletion of the 110-kilodalton isoform of poly(ADP-ribose) glycohydrolase
increases sensitivity to genotoxic and endotoxic stress in mice. Molecular and Cellular Biol-
ogy, 24(16), 7163–7178. https://doi.org/10.1128/MCB.24.16.7163-7178.2004.
Croteau, D. L., Fang, E. F., Nilsen, H., & Bohr, V. A. (2017). NAD(+) in DNA repair and
mitochondrial maintenance. Cell Cycle (Georgetown, TX), 16, 491–492. https://doi.org/
10.1080/15384101.2017.1285631.
Davis, A. A., Andruska, K. M., Benitez, B. A., Racette, B. A., Perlmutter, J. S., &
Cruchaga, C. (2016). Variants in GBA, SNCA, and MAPT influence Parkinson disease
136 James B. Kirkland and Mirella L. Meyer-Ficca

risk, age at onset, and progression. Neurobiology of Aging, 37, 209.e1–209.e7. https://doi.
org/10.1016/j.neurobiolaging.2015.09.014.
deGoma, E. M., Salavati, A., Shinohara, R. T., Saboury, B., Pollan, L., Schoen, M., et al.
(2015). A pilot trial to examine the effect of high-dose niacin on arterial wall inflamma-
tion using fluorodeoxyglucose positron emission tomography. Academic Radiology, 22(5),
600–609. https://doi.org/10.1016/j.acra.2014.12.015.
Deibel, S. H., Zelinski, E. L., Keeley, R. J., Kovalchuk, O., & McDonald, R. J. (2015). Epi-
genetic alterations in the suprachiasmatic nucleus and hippocampus contribute to age-
related cognitive decline. Oncotarget, 6(27), 23181–23203. https://doi.org/10.18632/
oncotarget.4036.
Durnin, L., Dai, Y., Aiba, I., Shuttleworth, C. W., Yamboliev, I. A., & Mutafova-
Yambolieva, V. N. (2012). Release, neuronal effects and removal of extracellular
β-nicotinamide adenine dinucleotide (β-NAD+) in the rat brain. The European Journal
of Neuroscience, 35(3), 423–435. https://doi.org/10.1111/j.1460-9568.2011.07957.x.
Elvehjem, C. A., Madden, R. J., Strong, F. M., & Woolley, D. W. (1937). Relation of nic-
otinic acid and nicotinic acid amide to canine black tongue. Journal of the American Chem-
ical Society, 59(9), 1767–1768. https://doi.org/10.1021/ja01288a509.
Emanuelli, M., Carnevali, F., Saccucci, F., Pierella, F., Amici, A., Raffaelli, N., et al. (2001).
Molecular cloning, chromosomal localization, tissue mRNA levels, bacterial expression,
and enzymatic properties of human NMN adenylyltransferase. The Journal of Biological
Chemistry, 276(1), 406–412. https://doi.org/10.1074/jbc.M008700200.
Escande, C., Nin, V., Price, N. L., Capellini, V., Gomes, A. P., Barbosa, M. T., et al. (2013).
Flavonoid apigenin is an inhibitor of the NAD + ase CD38: Implications for cellular
NAD+ metabolism, protein acetylation, and treatment of metabolic syndrome.
Diabetes, 62(4), 1084–1093. https://doi.org/10.2337/db12-1139.
Etheridge, E. W. (1972). The butterfly caste: A social history of pellagra in the South. Westport,
CT: Greenwood Publishing Co.
Fang, E. F., & Bohr, V. A. (2017). NAD(+): The convergence of DNA repair and
mitophagy. Autophagy, 13(2), 442–443. https://doi.org/10.1080/15548627.2016.
1257467.
Fang, E. F., Kassahun, H., Croteau, D. L., Scheibye-Knudsen, M., Marosi, K., Lu, H., et al.
(2016). NAD(+) replenishment improves lifespan and healthspan in ataxia telangiectasia
models via mitophagy and DNA repair. Cell Metabolism, 24(4), 566–581. https://doi.
org/10.1016/j.cmet.2016.09.004.
Fang, E. F., Scheibye-Knudsen, M., Brace, L. E., Kassahun, H., SenGupta, T., Nilsen, H.,
et al. (2014). Defective mitophagy in XPA via PARP-1 hyperactivation and NAD
(+)/SIRT1 reduction. Cell, 157(4), 882–896. https://doi.org/10.1016/j.cell.2014.03.
026.
Fazio, S., Guyton, J. R., Lin, J., Tomassini, J. E., Shah, A., & Tershakovec, A. M. (2010).
Long-term efficacy and safety of ezetimibe/simvastatin coadministered with extended-
release niacin in hyperlipidaemic patients with diabetes or metabolic syndrome. Diabetes,
Obesity & Metabolism, 12(11), 983–993. https://doi.org/10.1111/j.1463-1326.2010.
01289.x.
Feldblum, S., Rougier, A., Loiseau, H., Loiseau, P., Cohadon, F., Morselli, P. L., et al.
(1988). Quinolinic-phosphoribosyl transferase activity is decreased in epileptic human
brain tissue. Epilepsia, 29(5), 523–529.
Ferrero, E., Lo Buono, N., Horenstein, A. L., Funaro, A., & Malavasi, F. (2014). The ADP-
ribosyl cyclases—The current evolutionary state of the ARCs. Frontiers in Bioscience
(Landmark Edition), 19, 986–1002.
Foster, A. C., Whetsell, W. O., Bird, E. D., & Schwarcz, R. (1985). Quinolinic acid phos-
phoribosyltransferase in human and rat brain: Activity in Huntington’s disease and in
quinolinate-lesioned rat striatum. Brain Research, 336(2), 207–214.
Niacin 137

Fouquerel, E., & Sobol, R. W. (2014). ARTD1 (PARP1) activation and NAD+ in DNA
repair and cell death. DNA Repair, 23, 27–32. https://doi.org/10.1016/j.dnarep.2014.
09.004.
Frederick, D. W., Davis, J. G., Dávila, A., Agarwal, B., Michan, S., Puchowicz, M. A., et al.
(2015). Increasing NAD synthesis in muscle via nicotinamide phosphoribosyltransferase
is not sufficient to promote oxidative metabolism. Journal of Biological Chemistry, 290(3),
1546–1558. https://doi.org/10.1074/jbc.M114.579565.
Fu, C. S., Swendseid, M. E., Jacob, R. A., & McKee, R. W. (1989). Biochemical markers for
assessment of niacin status in young men: Levels of erythrocyte niacin coenzymes and
plasma tryptophan. The Journal of Nutrition, 119(12), 1949–1955.
Fukuwatari, T., Ohta, M., Kimtjra, N., Sasaki, R., & Shibata, K. (2004). Conversion ratio of
tryptophan to niacin in Japanese women fed a purified diet conforming to the Japanese
Dietary Reference Intakes. Journal of Nutritional Science and Vitaminology, 50(6), 385–391.
Fukuwatari, T., Sugimoto, E., & Shibata, K. (2002). Growth-promoting activity of
pyrazinoic acid, a putative active compound of antituberculosis drug pyrazinamide, in
niacin-deficient rats through the inhibition of ACMSD activity. Bioscience, Biotechnology,
and Biochemistry, 66(7), 1435–1441. https://doi.org/10.1271/bbb.66.1435.
Ganji, S. H., Qin, S., Zhang, L., Kamanna, V. S., & Kashyap, M. L. (2009). Niacin inhibits
vascular oxidative stress, redox-sensitive genes, and monocyte adhesion to human aortic
endothelial cells. Atherosclerosis, 202(1), 68–75. https://doi.org/10.1016/j.
atherosclerosis.2008.04.044.
Gaziano, J. M., & Manson, J. E. (1996). Diet and heart disease. The role of fat, alcohol, and
antioxidants. Cardiology Clinics, 14(1), 69–83.
Gehring, W. (2004). Nicotinic acid/niacinamide and the skin. Journal of Cosmetic Dermatology,
3(2), 88–93. https://doi.org/10.1111/j.1473-2130.2004.00115.x.
Gensler, H. L., Williams, T., Huang, A. C., & Jacobson, E. L. (1999). Oral niacin prevents
photocarcinogenesis and photoimmunosuppression in mice. Nutrition and Cancer, 34(1),
36–41. https://doi.org/10.1207/S15327914NC340105.
Goldie, C., Taylor, A. J., Nguyen, P., McCoy, C., Zhao, X.-Q., & Preiss, D. (2015). Niacin
therapy and the risk of new-onset diabetes: A meta-analysis of randomised controlled
trials. Heart (British Cardiac Society), 102, 198–203. https://doi.org/10.1136/heartjnl-
2015-308055.
Gopalan, C. (1969). Possible role for dietary leucine in the pathogenesis of pellagra. The
Lancet (London, England), 1(7587), 197–199.
Gopalan, C. (2009). Leucine and pellagra. Nutrition Reviews, 26(11), 323–326. https://doi.
org/10.1111/j.1753-4887.1968.tb00835.x.
Gordon, T., Castelli, W. P., Hjortland, M. C., Kannel, W. B., & Dawber, T. R. (1977).
High density lipoprotein as a protective factor against coronary heart disease. The Fra-
mingham Study. The American Journal of Medicine, 62(5), 707–714.
Goudarzi, A., Zhang, D., Huang, H., Barral, S., Kwon, O. K., Qi, S., et al. (2016). Dynamic
competing histone H4 K5K8 acetylation and butyrylation are hallmarks of highly active
gene promoters. Molecular Cell, 62(2), 169–180. https://doi.org/10.1016/j.molcel.2016.
03.014.
Grabowska, W., Sikora, E., & Bielak-Zmijewska, A. (2017). Sirtuins, a promising target in
slowing down the ageing process. Biogerontology, 18, 447–476. https://doi.org/10.1007/
s10522-017-9685-9.
Grant, R., & Kapoor, V. (2003). Inhibition of indoleamine 2,3-dioxygenase activity in IFN-
gamma stimulated astroglioma cells decreases intracellular NAD levels. Biochemical Phar-
macology, 66(6), 1033–1036.
Grube, K., & B€ urkle, A. (1992). Poly(ADP-ribose) polymerase activity in mononuclear leu-
kocytes of 13 mammalian species correlates with species-specific life span. Proceedings of
the National Academy of Sciences of the United States of America, 89(24), 11759–11763.
138 James B. Kirkland and Mirella L. Meyer-Ficca

Gupte, R., Liu, Z., & Kraus, W. L. (2017). PARPs and ADP-ribosylation: Recent advances
linking molecular functions to biological outcomes. Genes & Development, 31(2),
101–126. https://doi.org/10.1101/gad.291518.116.
Guse, A. H., & Lee, H. C. (2008). NAADP: A universal Ca2 + trigger. Science Signaling,
1(44), re10. https://doi.org/10.1126/scisignal.144re10.
Guyton, J. R., Fazio, S., Adewale, A. J., Jensen, E., Tomassini, J. E., Shah, A., et al. (2012).
Effect of extended-release niacin on new-onset diabetes among hyperlipidemic patients
treated with ezetimibe/simvastatin in a randomized controlled trial. Diabetes Care, 35(4),
857–860. https://doi.org/10.2337/dc11-1369.
Guyton, J. R., Goldberg, A. C., Kreisberg, R. A., Sprecher, D. L., Superko, H. R., &
O’Connor, C. M. (1998). Effectiveness of once-nightly dosing of extended-release nia-
cin alone and in combination for hypercholesterolemia. The American Journal of Cardiol-
ogy, 82(6), 737–743.
Hegyi, J., Schwartz, R. A., & Hegyi, V. (2004). Pellagra: Dermatitis, dementia, and diarrhea.
International Journal of Dermatology, 43(1), 1–5.
Herceg, Z., & Wang, Z.-Q. (2001). Functions of poly(ADP-ribose) polymerase (PARP) in
DNA repair, genomic integrity and cell death. Mutation Research/Fundamental and
Molecular Mechanisms of Mutagenesis, 477(1–2), 97–110. https://doi.org/10.1016/
S0027-5107(01)00111-7.
Hochholzer, W., Berg, D. D., & Giugliano, R. P. (2011). The facts behind niacin. Therapeutic
Advances in Cardiovascular Disease, 5(5), 227–240. https://doi.org/10.1177/
1753944711419197.
Horwitt, M. K., Harper, A. E., & Henderson, L. M. (1981). Niacin-tryptophan relationships
for evaluating niacin equivalents. The American Journal of Clinical Nutrition, 34(3),
423–427.
Hottiger, M. O., Hassa, P. O., Luscher, B., Schuler, H., & Koch-Nolte, F. (2010). Toward
a unified nomenclature for mammalian ADP-ribosyltransferases. Trends in Biochemical
Sciences, 35(4), 208–219. https://doi.org/10.1016/j.tibs.2009.12.003.
Howitz, K. T., Bitterman, K. J., Cohen, H. Y., Lamming, D. W., Lavu, S., Wood, J. G., et al.
(2003). Small molecule activators of sirtuins extend Saccharomyces cerevisiae lifespan.
Nature, 425(6954), 191–196. https://doi.org/10.1038/nature01960.
Hussein, A. A., & Nicholls, S. J. (2010). Critical appraisal of laropiprant and extended-release
niacin combination in the management of mixed dyslipidemias and primary hypercho-
lesterolemia. Therapeutics and Clinical Risk Management, 6, 183–190.
Imai, S. (2009a). From heterochromatin islands to the NAD World: A hierarchical view of
aging through the functions of mammalian Sirt1 and systemic NAD biosynthesis.
Biochimica et Biophysica Acta, 1790(10), 997–1004. https://doi.org/10.1016/j.bbagen.
2009.03.005.
Imai, S.-I. (2009b). The NAD world: A new systemic regulatory network for metabolism and
aging—Sirt1, systemic NAD biosynthesis, and their importance. Cell Biochemistry and
Biophysics, 53(2), 65–74. https://doi.org/10.1007/s12013-008-9041-4.
Imai, S.-I. (2010). “Clocks” in the NAD World: NAD as a metabolic oscillator for the reg-
ulation of metabolism and aging. Biochimica et Biophysica Acta, 1804(8), 1584–1590.
https://doi.org/10.1016/j.bbapap.2009.10.024.
Imai, S. (2011). Dissecting systemic control of metabolism and aging in the NAD World: The
importance of SIRT1 and NAMPT-mediated NAD biosynthesis. FEBS Letters, 585(11),
1657–1662. https://doi.org/10.1016/j.febslet.2011.04.060.
Imai, S., & Guarente, L. (2014). NAD + and sirtuins in aging and disease. Trends in Cell Biol-
ogy, 24(8), 464–471. https://doi.org/10.1016/j.tcb.2014.04.002.
Imai, S., & Guarente, L. (2016). It takes two to tango: NAD + and sirtuins in aging/longevity
control. NPJ Aging and Mechanisms of Disease, 2(1), 16017. https://doi.org/10.1038/
npjamd.2016.17.
Niacin 139

Institute of Medicine. (1998). Dietary reference intakes for thiamin, riboflavin, niacin, vitamin B₆,
folate, vitamin B₁₂, pantothenic acid, biotin, and choline. Washington, DC: The National
Academies Press.
Jackson, T. M., Rawling, J. M., Roebuck, B. D., & Kirkland, J. B. (1995). Large supplements
of nicotinic acid and nicotinamide increase tissue NAD + and poly(ADP-ribose) levels
but do not affect diethylnitrosamine-induced altered hepatic foci in Fischer-344 rats. The
Journal of Nutrition, 125(6), 1455–1461.
Jacobson, E. L. (1993). Niacin deficiency and cancer in women. Journal of the American College
of Nutrition, 12(4), 412–416.
Jacobson, E. L., & Jacobson, M. K. (1997). Tissue NAD as a biochemical measure of niacin
status in humans. Methods in Enzymology, 280, 221–230.
Jacobson, E. L., Kim, H., Kim, M., Williams, J. D., Coyle, D. L., Coyle, W. R., et al. (2007).
A topical lipophilic niacin derivative increases NAD, epidermal differentiation and bar-
rier function in photodamaged skin. Experimental Dermatology, 16(6), 490–499. https://
doi.org/10.1111/j.1600-0625.2007.00553.x.
Jacobson, E. L., Shieh, W. M., & Huang, A. C. (1999). Mapping the role of NAD metab-
olism in prevention and treatment of carcinogenesis. Molecular and Cellular Biochemistry,
193(1–2), 69–74.
Jagielska, G., Tomaszewicz-Libudzic, E. C., & Brzozowska, A. (2007). Pellagra: A rare com-
plication of anorexia nervosa. European Child & Adolescent Psychiatry, 16(7), 417–420.
https://doi.org/10.1007/s00787-007-0613-4.
Jasper, H. (2013). Sirtuins: Longevity focuses on NAD +. Nature Chemical Biology, 9(11),
666–667. https://doi.org/10.1038/nchembio.1369.
Jasperson, L. K., Bucher, C., Panoskaltsis-Mortari, A., Mellor, A. L., Munn, D. H., &
Blazar, B. R. (2009). Inducing the tryptophan catabolic pathway, indoleamine 2,3-
dioxygenase (IDO), for suppression of graft-versus-host disease (GVHD) lethality.
Blood, 114(24), 5062–5070. https://doi.org/10.1182/blood-2009-06-227587.
Jeggo, P. A. (1998). DNA repair: PARP—Another guardian angel? Current Biology: CB, 8(2),
R49–51.
Jia, H., Li, X., Gao, H., Feng, Z., Li, X., Zhao, L., et al. (2008). High doses of nicotinamide
prevent oxidative mitochondrial dysfunction in a cellular model and improve motor def-
icit in a Drosophila model of Parkinson’s disease. Journal of Neuroscience Research, 86(9),
2083–2090. https://doi.org/10.1002/jnr.21650.
Jiang, Y., Liu, J., Chen, D., Yan, L., & Zheng, W. (2017). Sirtuin inhibition: Strategies,
inhibitors, and therapeutic potential. Trends in Pharmacological Sciences, 38(5), 459–472.
https://doi.org/10.1016/j.tips.2017.01.009.
Johansson, M. (1999). A human poly(ADP-ribose) polymerase gene family (ADPRTL):
cDNA cloning of two novel poly(ADP-ribose) polymerase homologues. Genomics,
57(3), 442–445. https://doi.org/10.1006/geno.1999.5799.
Juarez-Salinas, H., Levi, V., Jacobson, E. L., & Jacobson, M. K. (1982). Poly(ADP-ribose) has
a branched structure in vivo. The Journal of Biological Chemistry, 257(2), 607–609.
Kane, J. P. (1990). Regression of coronary atherosclerosis during treatment of familial hyper-
cholesterolemia with combined drug regimens. JAMA: The Journal of the American Medical
Association, 264(23), 3007. https://doi.org/10.1001/jama.1990.03450230043027.
Karlberg, T., Klepsch, M., Thorsell, A.-G., Andersson, C. D., Linusson, A., & Sch€ uler, H.
(2015). Structural basis for lack of ADP-ribosyltransferase activity in poly(ADP-ribose)
polymerase-13/zinc finger antiviral protein. The Journal of Biological Chemistry, 290(12),
7336–7344. https://doi.org/10.1074/jbc.M114.630160.
Kasamatsu, A., Nakao, M., Smith, B. C., Comstock, L. R., Ono, T., Kato, J., et al. (2011).
Hydrolysis of O-acetyl-ADP-ribose isomers by ADP-ribosylhydrolase 3. The Journal of
Biological Chemistry, 286(24), 21110–21117. https://doi.org/10.1074/jbc.M111.
237636.
140 James B. Kirkland and Mirella L. Meyer-Ficca

Kennedy, B. E., Sharif, T., Martell, E., Dai, C., Kim, Y., Lee, P. W. K., et al. (2016). NAD
(+) salvage pathway in cancer metabolism and therapy. Pharmacological Research, 114,
274–283. https://doi.org/10.1016/j.phrs.2016.10.027.
Kiehlbauch, C. C., Aboul-Ela, N., Jacobson, E. L., Ringer, D. P., & Jacobson, M. K. (1993).
High resolution fractionation and characterization of ADP-ribose polymers. Analytical
Biochemistry, 208(1), 26–34.
Kim, S.-Y., Cohen, B. M., Chen, X., Lukas, S. E., Shinn, A. K., Yuksel, A. C., et al. (2017).
Redox dysregulation in schizophrenia revealed by in vivo NAD +/NADH measure-
ment. Schizophrenia Bulletin, 43(1), 197–204. https://doi.org/10.1093/schbul/sbw129.
Klaidman, L., Morales, M., Kem, S., Yang, J., Chang, M.-L., & Adams, J. D. (2003). Nic-
otinamide offers multiple protective mechanisms in stroke as a precursor for NAD+, as a
PARP inhibitor and by partial restoration of mitochondrial function. Pharmacology,
69(3), 150–157. https://doi.org/72668.
Koehn, C. J., Jr., & Elvehjem, C. A. (1937). Further studies on the concentration of the anti-
pellagra factor. Journal of Biological Chemistry, 118, 693–699.
Koh, D. W., Lawler, A. M., Poitras, M. F., Sasaki, M., Wattler, S., Nehls, M. C., et al.
(2004). Failure to degrade poly(ADP-ribose) causes increased sensitivity to cytotoxicity
and early embryonic lethality. Proceedings of the National Academy of Sciences of the United
States of America, 101(51), 17699–17704. https://doi.org/10.1073/pnas.0406182101.
Kraus, W. L. (2015). PARPs and ADP-ribosylation: 50 years … and counting. Molecular Cell,
58(6), 902–910. https://doi.org/10.1016/j.molcel.2015.06.006.
Kraut, J. A., & Sachs, G. (2005). Hartnup disorder: Unraveling the mystery. Trends in Phar-
macological Sciences, 26(2), 53–55. https://doi.org/10.1016/j.tips.2004.12.003.
Kritikou, E. (2008). Circadian genetics: An enzymatic rheostat. Nature Reviews Genetics, 9(9),
653.
Kudo, Y., & Boyd, C. A. (2000). Human placental indoleamine 2,3-dioxygenase: Cellular
localization and characterization of an enzyme preventing fetal rejection. Biochimica et
Biophysica Acta, 1500(1), 119–124.
Kumar, V., & Takahashi, J. S. (2010). PARP around the clock. Cell, 142(6), 841–843.
https://doi.org/10.1016/j.cell.2010.08.037.
Kupper, J. H., van Gool, L., & Burkle, A. (1995). Molecular genetic systems to study the role
of poly(ADP-ribosyl)ation in the cellular response to DNA damage. Biochimie, 77(6),
450–455.
Lai, E., De Lepeleire, I., Crumley, T. M., Liu, F., Wenning, L. A., Michiels, N., et al. (2007).
Suppression of niacin-induced vasodilation with an antagonist to prostaglandin d2 recep-
tor subtype 1. Clinical Pharmacology and Therapeutics, 81(6), 849–857. https://doi.org/10.
1038/sj.clpt.6100180.
Lamon-Fava, S., Diffenderfer, M. R., Barrett, P. H. R., Buchsbaum, A., Nyaku, M.,
Horvath, K. V., et al. (2008). Extended-release niacin alters the metabolism of plasma apo-
lipoprotein (Apo) A-I and ApoB-containing lipoproteins. Arteriosclerosis, Thrombosis, and
Vascular Biology, 28(9), 1672–1678. https://doi.org/10.1161/ATVBAHA.108.164541.
la Paz, S. M., Bermudez, B., Naranjo, M. C., Lopez, S., Abia, R., & Muriana, F. J. G. (2016).
Pharmacological effects of niacin on acute hyperlipemia. Current Medicinal Chemistry,
23(25), 2826–2835.
Lee, J. M. S., Robson, M. D., Yu, L.-M., Shirodaria, C. C., Cunnington, C., Kylintireas, I.,
et al. (2009). Effects of high-dose modified-release nicotinic acid on atherosclerosis and
vascular function: A randomized, placebo-controlled, magnetic resonance imaging
study. Journal of the American College of Cardiology, 54(19), 1787–1794. https://doi.org/
10.1016/j.jacc.2009.06.036.
Lehmann, S., Loh, S. H. Y., & Martins, L. M. (2017). Enhancing NAD(+) salvage metab-
olism is neuroprotective in a PINK1 model of Parkinson’s disease. Biology Open, 6(2),
141–147. https://doi.org/10.1242/bio.022186.
Niacin 141

Li, F., Chong, Z. Z., & Maiese, K. (2006). Cell life versus cell longevity: The mysteries sur-
rounding the NAD + precursor nicotinamide. Current Medicinal Chemistry, 13(8),
883–895.
Li, R., Yu, K., Wang, Q., Wang, L., Mao, J., & Qian, J. (2016). Pellagra secondary to med-
ication and alcoholism: A case report and review of the literature. Nutrition in Clinical
Practice: Official Publication of the American Society for Parenteral and Enteral Nutrition,
31(6), 785–789. https://doi.org/10.1177/0884533616660991.
Liberti, M. V., & Locasale, J. W. (2016). The Warburg effect: How does it benefit cancer
cells? Trends in Biochemical Sciences, 41(3), 211–218. https://doi.org/10.1016/j.tibs.
2015.12.001.
Lin, S.-J., Kaeberlein, M., Andalis, A. A., Sturtz, L. A., Defossez, P.-A., Culotta, V. C., et al.
(2002). Calorie restriction extends Saccharomyces cerevisiae lifespan by increasing res-
piration. Nature, 418(6895), 344–348. https://doi.org/10.1038/nature00829.
Lloyd-Jones, D. M. (2014). Niacin and HDL cholesterol—Time to face facts. New England
Journal of Medicine, 371(3), 271–273. https://doi.org/10.1056/NEJMe1406410.
Lord, C. J., & Ashworth, A. (2017). PARP inhibitors: Synthetic lethality in the clinic. Science
(New York, NY), 355(6330), 1152–1158. https://doi.org/10.1126/science.aam7344.
Magni, G., Amici, A., Emanuelli, M., Orsomando, G., Raffaelli, N., & Ruggieri, S. (2004).
Enzymology of NAD+ homeostasis in man. Cellular and Molecular Life Sciences: CMLS,
61(1), 19–34. https://doi.org/10.1007/s00018-003-3161-1.
Maltos, A. L., Portari, G. V., Moraes, G. V., Monteiro, M. C. R., Vannucchi, H., & da
Cunha, D. F. (2015). Niacin metabolism and indoleamine 2,3-dioxygenase activation
in malnourished patients with flaky paint dermatosis. Nutrition (Burbank, Los Angeles
County, CA), 31(6), 890–892. https://doi.org/10.1016/j.nut.2014.12.023.
Mashimo, M., Kato, J., & Moss, J. (2013). ADP-ribosyl-acceptor hydrolase 3 regulates poly
(ADP-ribose) degradation and cell death during oxidative stress. Proceedings of the National
Academy of Sciences of the United States of America, 110(47), 18964–18969. https://doi.org/
10.1073/pnas.1312783110.
Mashimo, M., Kato, J., & Moss, J. (2014). Structure and function of the ARH family of
ADP-ribosyl-acceptor hydrolases. DNA Repair, 23, 88–94. https://doi.org/10.1016/j.
dnarep.2014.03.005.
Massudi, H., Grant, R., Braidy, N., Guest, J., Farnsworth, B., & Guillemin, G. J. (2012).
Age-associated changes in oxidative stress and NAD+ metabolism in human tissue. PLoS
One, 7(7). e42357 https://doi.org/10.1371/journal.pone.0042357.
Masutani, M., & Fujimori, H. (2013). Poly(ADP-ribosyl)ation in carcinogenesis. Molecular
Aspects of Medicine, 34(6), 1202–1216. https://doi.org/10.1016/j.mam.2013.05.003.
Masutani, M., Nakagama, H., & Sugimura, T. (2003). Poly(ADP-ribose) and carcinogenesis.
Genes, Chromosomes & Cancer, 38(4), 339–348. https://doi.org/10.1002/gcc.10250.
Masutani, M., Nozaki, T., Nakamoto, K., Nakagama, H., Suzuki, H., Kusuoka, O., et al.
(2000). The response of Parp knockout mice against DNA damaging agents. Mutation
Research, 462(2–3), 159–166.
Mazucanti, C. H., Cabral-Costa, J. V., Vasconcelos, A. R., Andreotti, D. Z., Scavone, C., &
Kawamoto, E. M. (2015). Longevity pathways (mTOR, SIRT, insulin/IGF-1) as key
modulatory targets on aging and neurodegeneration. Current Topics in Medicinal Chemis-
try, 15(21), 2116–2138.
McKenney, J. M., Jones, P. H., Bays, H. E., Knopp, R. H., Kashyap, M. L., Ruoff, G. E.,
et al. (2007). Comparative effects on lipid levels of combination therapy with a statin and
extended-release niacin or ezetimibe versus a statin alone (the COMPELL study).
Atherosclerosis, 192(2), 432–437. https://doi.org/10.1016/j.atherosclerosis.2006.11.037.
Meyer, R. G., Meyer-Ficca, M. L., Jacobson, E. L., & Jacobson, M. K. (2004). Enzymes in
poly(ADP-ribose) metabolism. In A. Burkle (Ed.), Book section: Vol. 1. Poly(ADP-ribosyl)
ation: Landes Bioscience.
142 James B. Kirkland and Mirella L. Meyer-Ficca

Meyer, R. G., Meyer-Ficca, M. L., Whatcott, C. J., Jacobson, E. L., & Jacobson, M. K.
(2007). Two small enzyme isoforms mediate mammalian mitochondrial poly(ADP-
ribose) glycohydrolase (PARG) activity. Experimental Cell Research, 313(13),
2920–2936https://doi.org/10.1016/j.yexcr.2007.03.043.
Meyer, R., Muller, M., Beneke, S., Kupper, J. H., & Burkle, A. (2000). Negative regulation
of alkylation-induced sister-chromatid exchange by poly(ADP-ribose) polymerase-1
activity. International Journal of Cancer, 88(3), 351–355. https://doi.org/10.1002/1097-
0215(20001101)88:3<351::AID-IJC5>3.0.CO;2-H.
Meyer-Ficca, M. L., Ihara, M., Bader, J. J., Leu, N. A., Beneke, S., & Meyer, R. G. (2015).
Spermatid head elongation with normal nuclear shaping requires ADP-ribosyltransferase
PARP11 (ARTD11) in mice. Biology of Reproduction, 92(3), 80. https://doi.org/10.
1095/biolreprod.114.123661.
Meyer-Ficca, M. L., Meyer, R. G., Coyle, D. L., Jacobson, E. L., & Jacobson, M. K. (2004).
Human poly(ADP-ribose) glycohydrolase is expressed in alternative splice variants yield-
ing isoforms that localize to different cell compartments. Experimental Cell Research,
297(2), 521–532. https://doi.org/10.1016/j.yexcr.2004.03.050.
Meyer-Ficca, M. L., Meyer, R. G., Jacobson, E. L., & Jacobson, M. K. (2005). Poly(ADP-
ribose) polymerases: Managing genome stability. The International Journal of Biochemistry &
Cell Biology, 37(5), 920–926. https://doi.org/10.1016/j.biocel.2004.09.011.
Michan, S., & Sinclair, D. (2007). Sirtuins in mammals: Insights into their biological func-
tion. The Biochemical Journal, 404(1), 1–13. https://doi.org/10.1042/BJ20070140.
Miller, D. F. (1978). Pellagra deaths in the United States. The American Journal of Clinical Nutri-
tion, 31(4), 558–559.
Miwa, M., Ishihara, M., Takishima, S., Takasuka, N., Maeda, M., Yamaizumi, Z., et al.
(1981). The branching and linear portions of poly(adenosine diphosphate ribose) have
the same alpha(1 leads to 2) ribose-ribose linkage. Journal of Biological Chemistry,
256(6), 2916–2921.
Miwa, M., Saikawa, N., Yamaizumi, Z., Nishimura, S., & Sugimura, T. (1979). Structure
of poly(adenosine diphosphate ribose): Identification of 20 -[100 -ribosyl-200 -(or 300 -)(100 -
ribosyl)]adenosine-50 ,500 ,500 -tris(phosphate) as a branch linkage. Proceedings of the National
Academy of Sciences of the United States of America, 76(2), 595–599.
Miwa, M., & Sugimura, T. (1984). Structure of poly(ADP-ribose). Methods in Enzymology,
106, 441–450.
Módis, K., Gero, D., Erdelyi, K., Szoleczky, P., DeWitt, D., & Szabo, C. (2012). Cellular
bioenergetics is regulated by PARP1 under resting conditions and during oxidative
stress. Biochemical Pharmacology, 83(5), 633–643. https://doi.org/10.1016/j.bcp.2011.
12.014.
Mouchiroud, L., Houtkooper, R. H., Moullan, N., Katsyuba, E., Ryu, D., Cantó, C., et al.
(2013). The NAD(+)/sirtuin pathway modulates longevity through activation of mito-
chondrial UPR and FOXO signaling. Cell, 154(2), 430–441. https://doi.org/10.1016/j.
cell.2013.06.016.
Muller, A. J., & Prendergast, G. C. (2007). Indoleamine 2,3-dioxygenase in immune sup-
pression and cancer. Current Cancer Drug Targets, 7(1), 31–40.
Munn, D. H., & Mellor, A. L. (2013). Indoleamine 2,3 dioxygenase and metabolic control of
immune responses. Trends in Immunology, 34(3), 137–143. https://doi.org/10.1016/j.it.
2012.10.001.
Munn, D. H., & Mellor, A. L. (2016). IDO in the tumor microenvironment: Inflammation,
counter-regulation, and tolerance. Trends in Immunology, 37(3), 193–207. https://doi.
org/10.1016/j.it.2016.01.002.
Murray, M. F., Langan, M., & MacGregor, R. R. (2001). Increased plasma tryptophan in
HIV-infected patients treated with pharmacologic doses of nicotinamide. Nutrition
(Burbank, Los Angeles County, CA), 17(7–8), 654–656.
Niacin 143

Mutafova-Yambolieva, V. N. (2012). Neuronal and extraneuronal release of ATP and NAD


(+) in smooth muscle. IUBMB Life, 64(10), 817–824. https://doi.org/10.1002/iub.
1076.
Mutafova-Yambolieva, V. N., Hwang, S. J., Hao, X., Chen, H., Zhu, M. X., Wood, J. D.,
et al. (2007). Beta-nicotinamide adenine dinucleotide is an inhibitory neurotransmitter
in visceral smooth muscle. Proceedings of the National Academy of Sciences of the United States
of America, 104(41), 16359–16364. https://doi.org/10.1073/pnas.0705510104.
Nakahata, Y., Kaluzova, M., Grimaldi, B., Sahar, S., Hirayama, J., Chen, D., et al. (2008).
The NAD +dependent deacetylase SIRT1 modulates CLOCK-mediated chromatin
remodeling and circadian control. Cell, 134(2), 329–340. https://doi.org/10.1016/j.
cell.2008.07.002.
Nakahata, Y., Sahar, S., Astarita, G., Kaluzova, M., & Sassone-Corsi, P. (2009). Circadian
control of the NAD + salvage pathway by CLOCK-SIRT1. Science (New York, NY),
324(5927), 654–657. https://doi.org/10.1126/science.1170803.
Nimmagadda, V. K., Bever, C. T., Vattikunta, N. R., Talat, S., Ahmad, V., Nagalla, N. K.,
et al. (2013). Overexpression of SIRT1 protein in neurons protects against experimental
autoimmune encephalomyelitis through activation of multiple SIRT1 targets. Journal of
Immunology (Baltimore, Md: 1950), 190(9), 4595–4607. https://doi.org/10.4049/
jimmunol.1202584.
Nimmagadda, V. K. C., Makar, T. K., Chandrasekaran, K., Sagi, A. R., Ray, J.,
Russell, J. W., et al. (2017). SIRT1 and NAD + precursors: Therapeutic targets in mul-
tiple sclerosis a review. Journal of Neuroimmunology, 304, 29–34. https://doi.org/10.
1016/j.jneuroim.2016.07.007.
Nogueira, A., Duarte, A. F., Magina, S., & Azevedo, F. (2009). Pellagra associated with
esophageal carcinoma and alcoholism. Dermatology Online Journal, 15(5), 8.
Nozaki, J., Dakeishi, M., Ohura, T., Inoue, K., Manabe, M., Wada, Y., et al. (2001). Homo-
zygosity mapping to chromosome 5p15 of a gene responsible for Hartnup disorder. Bio-
chemical and Biophysical Research Communications, 284(2), 255–260. https://doi.org/10.
1006/bbrc.2001.4961.
Ondracek, C. R., Frappier, V., Ringel, A. E., Wolberger, C., & Guarente, L. (2017). Muta-
tions that allow SIR2 orthologs to function in a NAD(+)-depleted environment. Cell
Reports, 18(10), 2310–2319. https://doi.org/10.1016/j.celrep.2017.02.031.
Ooi, E. M., Watts, G. F., Chan, D. C., Pang, J., Tenneti, V. S., Hamilton, S. J., et al. (2015).
Effects of extended-release niacin on the postprandial metabolism of Lp(a) and ApoB-
100-containing lipoproteins in statin-treated men with type 2 diabetes mellitus. Arterio-
sclerosis, Thrombosis, and Vascular Biology, 35(12), 2686–2693. https://doi.org/10.1161/
ATVBAHA.115.306136.
Opitz, C. A., & Heiland, I. (2015). Dynamics of NAD-metabolism: Everything but
constant. Biochemical Society Transactions, 43(6), 1127–1132. https://doi.org/10.1042/
BST20150133.
Palazzo, L., Mikoč, A., & Ahel, I. (2017). ADP-ribosylation: New facets of an ancient mod-
ification. The FEBS Journal, 284, 2932–2946. https://doi.org/10.1111/febs.14078.
Pan, H., & Finkel, T. (2017). Key proteins and pathways that regulate lifespan. The Journal of
Biological Chemistry, 292(16), 6452–6460. https://doi.org/10.1074/jbc.R116.771915.
Pang, J., Chan, D. C., Hamilton, S. J., Tenneti, V. S., Watts, G. F., & Barrett, P. H. R.
(2014). Effect of niacin on high-density lipoprotein apolipoprotein A-I kinetics in
statin-treated patients with type 2 diabetes mellitus. Arteriosclerosis, Thrombosis, and Vas-
cular Biology, 34(2), 427–432. https://doi.org/10.1161/ATVBAHA.113.302019.
Park, J. Y., Jang, S. Y., Shin, Y. K., Koh, H., Suh, D. J., Shinji, T., et al. (2013). Mito-
chondrial swelling and microtubule depolymerization are associated with energy deple-
tion in axon degeneration. Neuroscience, 238, 258–269. https://doi.org/10.1016/
j.neuroscience.2013.02.033.
144 James B. Kirkland and Mirella L. Meyer-Ficca

Park, S. M., Li, T., Wu, S., Li, W.-Q., Weinstock, M., Qureshi, A. A., et al. (2017). Niacin
intake and risk of skin cancer in US women and men. International Journal of Cancer,
140(9), 2023–2031. https://doi.org/10.1002/ijc.30630.
Partida-Sánchez, S., Cockayne, D. A., Monard, S., Jacobson, E. L., Oppenheimer, N.,
Garvy, B., et al. (2001). Cyclic ADP-ribose production by CD38 regulates intracellular
calcium release, extracellular calcium influx and chemotaxis in neutrophils and is
required for bacterial clearance in vivo. Nature Medicine, 7(11), 1209–1216. https://
doi.org/10.1038/nm1101-1209.
Partida-Sanchez, S., Gasser, A., Fliegert, R., Siebrands, C. C., Dammermann, W., Shi, G.,
et al. (2007). Chemotaxis of mouse bone marrow neutrophils and dendritic cells is con-
trolled by ADP-ribose, the major product generated by the CD38 enzyme reaction. Jour-
nal of Immunology (Baltimore, Md: 1950), 179(11), 7827–7839.
Patterson, J. I., Brown, R. R., Linkswiler, H., & Harper, A. E. (1980). Excretion of
tryptophan-niacin metabolites by young men: Effects of tryptophan, leucine, and vita-
min B6 intakes. The American Journal of Clinical Nutrition, 33(10), 2157–2167.
Peek, C. B., Affinati, A. H., Ramsey, K. M., Kuo, H.-Y., Yu, W., Sena, L. A., et al. (2013).
Circadian clock NAD+ cycle drives mitochondrial oxidative metabolism in mice. Science
(New York, NY), 342(6158), 1243417. https://doi.org/10.1126/science.1243417.
Penberthy, W. T. (2007). Pharmacological targeting of IDO-mediated tolerance for treating
autoimmune disease. Current Drug Metabolism, 8(3), 245–266.
Penberthy, W. T., & Kirkland, J. B. (2012). Niacin. In J. W. Erdman, I. A. Macdonald, &
S. H. Zeisel (Eds.), Present knowledge in nutrition (pp. 293–306). Oxford, UK:
Wiley-Blackwell. https://doi.org/10.1002/9781119946045.ch19.

Perina, D., Mikoč, A., Ahel, J., Cetkovi c, H., Žaja, R., & Ahel, I. (2014). Distribution of
protein poly(ADP-ribosyl)ation systems across all domains of life. DNA Repair, 23, 4–16.
https://doi.org/10.1016/j.dnarep.2014.05.003.
Pollak, N., D€ olle, C., & Ziegler, M. (2007). The power to reduce: Pyridine nucleotides—
Small molecules with a multitude of functions. Biochemical Journal, 402(2), 205–218.
https://doi.org/10.1042/BJ20061638.
Poltronieri, P., & Miwa, M. (2016). Editorial (thematic issue: Overview on ADP ribosylation
and PARP superfamily of proteins). Current Protein & Peptide Science, 17(7), 630–632.
https://doi.org/10.2174/138920371707160908172601.
Prousky, J., Millman, C. G., & Kirkland, J. B. (2011). Pharmacologic use of niacin. Journal of
Evidence-Based Complementary & Alternative Medicine, 16(2), 91–101. https://doi.org/10.
1177/2156587211399579.
Quarona, V., Zaccarello, G., Chillemi, A., Brunetti, E., Singh, V. K., Ferrero, E., et al.
(2013). CD38 and CD157: A long journey from activation markers to multifunctional
molecules. Cytometry Part B, Clinical Cytometry, 84(4), 207–217. https://doi.org/10.
1002/cyto.b.21092.
Raffaelli, N., Sorci, L., Amici, A., Emanuelli, M., Mazzola, F., & Magni, G. (2002). Iden-
tification of a novel human nicotinamide mononucleotide adenylyltransferase. Biochem-
ical and Biophysical Research Communications, 297(4), 835–840.
Rajakumar, K. (2000). Pellagra in the United States: A historical perspective. Southern Medical
Journal, 93(3), 272–277.
Rajawat, J., Shukla, N., & Mishra, D. P. (2017). Therapeutic targeting of poly(ADP-ribose)
polymerase-1 in cancer: Current developments, therapeutic strategies, and future oppor-
tunities. Medicinal Research Reviews, 37, 1461–1491. https://doi.org/10.1002/med.
21442.
Ramsey, K. M., Yoshino, J., Brace, C. S., Abrassart, D., Kobayashi, Y., Marcheva, B., et al.
(2009). Circadian clock feedback cycle through NAMPT-mediated NAD + biosynthe-
sis. Science (New York, NY), 324(5927), 651–654. https://doi.org/10.1126/science.
1171641.
Niacin 145

Rehan, L., Laszki-Szcza˛chor, K., Sobieszcza nska, M., & Polak-Jonkisz, D. (2014). SIRT1
and NAD as regulators of ageing. Life Sciences, 105(1–2), 1–6. https://doi.org/10.
1016/j.lfs.2014.03.015.
Revollo, J. R., Grimm, A. A., & Imai, S. (2004). The NAD biosynthesis pathway mediated
by nicotinamide phosphoribosyltransferase regulates Sir2 activity in mammalian cells.
The Journal of Biological Chemistry, 279(49), 50754–50763. https://doi.org/10.1074/
jbc.M408388200.
Revollo, J. R., Grimm, A. A., & Imai, S. (2007). The regulation of nicotinamide adenine
dinucleotide biosynthesis by Nampt/PBEF/visfatin in mammals. Current Opinion in Gas-
troenterology, 23(2), 164–170. https://doi.org/10.1097/MOG.0b013e32801b3c8f.
Rongvaux, A., Andris, F., Van Gool, F., & Leo, O. (2003). Reconstructing eukaryotic NAD
metabolism. BioEssays: News and Reviews in Molecular, Cellular and Developmental Biology,
25(7), 683–690. https://doi.org/10.1002/bies.10297.
Rubic, T., Trottmann, M., & Lorenz, R. L. (2004). Stimulation of CD36 and the key effec-
tor of reverse cholesterol transport ATP-binding cassette A1 in monocytoid cells by nia-
cin. Biochemical Pharmacology, 67(3), 411–419. https://doi.org/10.1016/j.bcp.2003.09.
014.
Satoh, A., Imai, S., & Guarente, L. (2017). The brain, sirtuins, and ageing. Nature Reviews
Neuroscience, 18(6), 362–374.
Scheibye-Knudsen, M., Mitchell, S. J., Fang, E. F., Iyama, T., Ward, T., Wang, J., et al.
(2014). A high-fat diet and NAD(+) activate Sirt1 to rescue premature aging in cockayne
syndrome. Cell Metabolism, 20(5), 840–855. https://doi.org/10.1016/j.cmet.2014.10.
005.
Schmeisser, K., Mansfeld, J., Kuhlow, D., Weimer, S., Priebe, S., Heiland, I., et al. (2013).
Role of sirtuins in lifespan regulation is linked to methylation of nicotinamide. Nature
Chemical Biology, 9(11), 693–700. https://doi.org/10.1038/nchembio.1352.
Schraufstatter, I. U., Hinshaw, D. B., Hyslop, P. A., Spragg, R. G., & Cochrane, C. G.
(1986). Oxidant injury of cells. DNA strand-breaks activate polyadenosine
diphosphate-ribose polymerase and lead to depletion of nicotinamide adenine dinucle-
otide. The Journal of Clinical Investigation, 77(4), 1312–1320. https://doi.org/10.1172/
JCI112436.
Schreiber, V., Hunting, D., Trucco, C., Gowans, B., Grunwald, D., De Murcia, G., et al.
(1995). A dominant-negative mutant of human poly(ADP-ribose) polymerase affects cell
recovery, apoptosis, and sister chromatid exchange following DNA damage. Proceedings
of the National Academy of Sciences of the United States of America, 92(11), 4753–4757.
Schuber, F., & Lund, F. E. (2004). Structure and enzymology of ADP-ribosyl cyclases: Con-
served enzymes that produce multiple calcium mobilizing metabolites. Current Molecular
Medicine, 4(3), 249–261.
Schuldt, A. (2010). Circadian rhythms: PARP1 feeds into clocks. Nature Reviews Molecular
Cell Biology, 11(11), 754–755. https://doi.org/10.1038/nrm2998.
Schultz, M. B., & Sinclair, D. A. (2016). Why NAD(+) declines during aging: It’s destroyed.
Cell Metabolism, 23(6), 965–966. https://doi.org/10.1016/j.cmet.2016.05.022.
Shah, G. M., Shah, R. G., Veillette, H., Kirkland, J. B., Pasieka, J. L., & Warner, R. R. P.
(2005). Biochemical assessment of niacin deficiency among carcinoid cancer patients.
The American Journal of Gastroenterology, 100(10), 2307–2314. https://doi.org/10.
1111/j.1572-0241.2005.00268.x.
Shall, S., & de Murcia, G. (2000). Poly(ADP-ribose) polymerase-1: What have we learned
from the deficient mouse model? Mutation Research, 460(1), 1–15.
Sharifi, R., Morra, R., Appel, C. D., Tallis, M., Chioza, B., Jankevicius, G., et al. (2013).
Deficiency of terminal ADP-ribose protein glycohydrolase TARG1/C6orf130 in neu-
rodegenerative disease. The EMBO Journal, 32(9), 1225–1237. https://doi.org/10.1038/
emboj.2013.51.
146 James B. Kirkland and Mirella L. Meyer-Ficca

Shibata, K., & Toda, S. (1997). Effects of sex hormones on the metabolism of tryptophan to
niacin and to serotonin in male rats. Bioscience, Biotechnology, and Biochemistry, 61(7),
1200–1202. https://doi.org/10.1271/bbb.61.1200.
Sibley, C. T., Vavere, A. L., Gottlieb, I., Cox, C., Matheson, M., Spooner, A., et al. (2013).
MRI-measured regression of carotid atherosclerosis induced by statins with and without
niacin in a randomised controlled trial: the NIA plaque study. Heart, 99(22), 1675–1680.
https://doi.org/10.1136/heartjnl-2013-303926.
Smith, J. S., Brachmann, C. B., Celic, I., Kenna, M. A., Muhammad, S., Starai, V. J., et al.
(2000). A phylogenetically conserved NAD+-dependent protein deacetylase activity in
the Sir2 protein family. Proceedings of the National Academy of Sciences of the United States of
America, 97(12), 6658–6663.
Smyth, L. M., Bobalova, J., Mendoza, M. G., Lew, C., & Mutafova-Yambolieva, V. N.
(2004). Release of beta-nicotinamide adenine dinucleotide upon stimulation of postgan-
glionic nerve terminals in blood vessels and urinary bladder. The Journal of Biological
Chemistry, 279(47), 48893–48903. https://doi.org/10.1074/jbc.M407266200.
Song, P., Ramprasath, T., Wang, H., & Zou, M.-H. (2017). Abnormal kynurenine pathway
of tryptophan catabolism in cardiovascular diseases. Cellular and Molecular Life Sciences:
CMLS, 74, 2899–2916. https://doi.org/10.1007/s00018-017-2504-2.
Srivastava, S. (2016). Emerging therapeutic roles for NAD(+) metabolism in mitochondrial
and age-related disorders. Clinical and Translational Medicine, 5(1), 25. https://doi.org/10.
1186/s40169-016-0104-7.
Surjana, D., Halliday, G. M., & Damian, D. L. (2010). Role of nicotinamide in DNA dam-
age, mutagenesis, and DNA repair. Journal of Nucleic Acids, 2010, 1–13. https://doi.org/
10.4061/2010/157591.
Tang, K., Sham, H., Hui, E., & Kirkland, J. B. (2008). Niacin deficiency causes oxidative
stress in rat bone marrow cells but not through decreased NADPH or glutathione status.
The Journal of Nutritional Biochemistry, 19(11), 746–753. https://doi.org/10.1016/j.
jnutbio.2007.10.003.
Tanner, K. G., Landry, J., Sternglanz, R., & Denu, J. M. (2000). Silent information regulator
2 family of NAD-dependent histone/protein deacetylases generates a unique product,
1-O-acetyl-ADP-ribose. Proceedings of the National Academy of Sciences of the United States
of America, 97(26), 14178–14182. https://doi.org/10.1073/pnas.250422697.
Taylor, A. J. (2004). Arterial biology for the investigation of the treatment effects of reducing
cholesterol (ARBITER) 2: A double-blind, placebo-controlled study of extended-
release niacin on atherosclerosis progression in secondary prevention patients treated
with statins. Circulation, 110(23), 3512–3517. https://doi.org/10.1161/01.CIR.
0000148955.19792.8D.
Taylor, A. J., Lee, H. J., & Sullenberger, L. E. (2006). The effect of 24 months of combi-
nation statin and extended-release niacin on carotid intima-media thickness: ARBITER
3. Current Medical Research and Opinion, 22(11), 2243–2250. https://doi.org/10.
1185/030079906X148508.
Taylor, A. J., Villines, T. C., Stanek, E. J., Devine, P. J., Griffen, L., Miller, M., et al. (2009).
Extended-release niacin or ezetimibe and carotid intima-media thickness. The New
England Journal of Medicine, 361(22), 2113–2122. https://doi.org/10.1056/
NEJMoa0907569.
Terada, N., Kinoshita, K., Taguchi, S., & Tokuda, Y. (2015). Wernicke encephalopathy and
pellagra in an alcoholic and malnourished patient. BMJ Case Reports, 2015, 1–3. https://
doi.org/10.1136/bcr-2015-209412.
The AIM-HIGH Investigators. (2011a). Niacin in patients with low HDL cholesterol levels
receiving intensive statin therapy. New England Journal of Medicine, 365(24), 2255–2267.
https://doi.org/10.1056/NEJMoa1107579.
Niacin 147

The AIM-HIGH Investigators. (2011b). The role of niacin in raising high-density lipoprotein
cholesterol to reduce cardiovascular events in patients with atherosclerotic cardiovascular
disease and optimally treated low-density lipoprotein cholesterol. American Heart Journal,
161(3), 471–477.e2. https://doi.org/10.1016/j.ahj.2010.11.017.
The AIM-HIGH Investigators. (2011c). The role of niacin in raising high-density lipoprotein
cholesterol to reduce cardiovascular events in patients with atherosclerotic cardiovascular
disease and optimally treated low-density lipoprotein cholesterol: Baseline characteristics
of study participants. The atherothrombosis intervention in metabolic syndrome with low HDL/
high triglycerides: Impact on global health outcomes (AIM-HIGH) trial, American Heart Journal,
161(3), 538–543. https://doi.org/10.1016/j.ahj.2010.12.007.
The HPS2-THRIVE Collaborative Group. (2014). Effects of extended-release niacin with
Laropiprant in high-risk patients. New England Journal of Medicine, 371(3), 203–212.
https://doi.org/10.1056/NEJMoa1300955.
Trucco, C., Rolli, V., Oliver, F. J., Flatter, E., Masson, M., Dantzer, F., et al. (1999). A dual
approach in the study of poly (ADP-ribose) polymerase: In vitro random mutagenesis
and generation of deficient mice. Molecular and Cellular Biochemistry, 193(1–2), 53–60.
van de Ven, R. A. H., Santos, D., & Haigis, M. C. (2017). Mitochondrial sirtuins and molec-
ular mechanisms of aging. Trends in Molecular Medicine, 23(4), 320–331. https://doi.org/
10.1016/j.molmed.2017.02.005.
van der Veer, E., Ho, C., O’Neil, C., Barbosa, N., Scott, R., Cregan, S. P., et al. (2007).
Extension of human cell lifespan by nicotinamide phosphoribosyltransferase. The Journal
of Biological Chemistry, 282(15), 10841–10845. https://doi.org/10.1074/jbc.
C700018200.
Van Gool, L., Meyer, R., Tobiasch, E., Cziepluch, C., Jauniaux, J. C., Mincheva, A., et al.
(1997). Overexpression of human poly(ADP-ribose) polymerase in transfected hamster
cells leads to increased poly(ADP-ribosyl)ation and cellular sensitization to gamma irra-
diation. European Journal of Biochemistry/FEBS, 244(1), 15–20.
Vazquez, B. N., Thackray, J. K., & Serrano, L. (2017). Sirtuins and DNA damage repair:
SIRT7 comes to play. Nucleus (Austin, TX), 8(2), 107–115. https://doi.org/10.
1080/19491034.2016.1264552.
Verdin, E. (2015). NAD+ in aging, metabolism, and neurodegeneration. Science, 350(6265),
1208–1213. https://doi.org/10.1126/science.aac4854.
Vergeer, M., Holleboom, A. G., Kastelein, J. J. P., & Kuivenhoven, J. A. (2010). The HDL
hypothesis: Does high-density lipoprotein protect from atherosclerosis? The Journal of
Lipid Research, 51(8), 2058–2073. https://doi.org/10.1194/jlr.R001610.
Vogt, A., Kassner, U., Hostalek, U., Peiter, A., & Steinhagen-Thiessen, E. (2006). Safety and
tolerability of nicotinic acid. Results of the multicenter, open, prospective NAUTILUS
study. MMW Fortschritte der Medizin, 148(18), 41.
Vogt, A., Kassner, U., Hostalek, U., Steinhagen-Thiessen, E., & NAUTILUS Study Group
(2006). Evaluation of the safety and tolerability of prolonged-release nicotinic acid in a
usual care setting: The NAUTILUS study. Current Medical Research and Opinion, 22(2),
417–425. https://doi.org/10.1185/030079906X89766.
Voight, B. F., Peloso, G. M., Orho-Melander, M., Frikke-Schmidt, R., Barbalic, M.,
Jensen, M. K., et al. (2012). Plasma HDL cholesterol and risk of myocardial infarction:
A Mendelian randomisation study. The Lancet (London, England), 380(9841), 572–580.
https://doi.org/10.1016/S0140-6736(12)60312-2.
Vyas, S., Matic, I., Uchima, L., Rood, J., Zaja, R., Hay, R. T., et al. (2014). Family-wide
analysis of poly(ADP-ribose) polymerase activity. Nature Communications, 5, 1–13.
https://doi.org/10.1038/ncomms5426.
Wakade, C., & Chong, R. (2014). A novel treatment target for Parkinson’s disease. Journal of
the Neurological Sciences, 347(1–2), 34–38. https://doi.org/10.1016/j.jns.2014.10.024.
148 James B. Kirkland and Mirella L. Meyer-Ficca

Wakade, C., Chong, R., Bradley, E., & Morgan, J. C. (2015). Low-dose niacin supplemen-
tation modulates GPR109A, niacin index and ameliorates Parkinson’s disease symptoms
without side effects. Clinical Case Reports, 3(7), 635–637. https://doi.org/10.1002/ccr3.
232.
Wakade, C., Chong, R., Bradley, E., Thomas, B., & Morgan, J. (2014). Upregulation of
GPR109A in Parkinson’s disease. PLoS One, 9(10) . e109818 https://doi.org/10.
1371/journal.pone.0109818.
Wang, Z. Q., Auer, B., Stingl, L., Berghammer, H., Haidacher, D., Schweiger, M., et al.
(1995). Mice lacking ADPRT and poly(ADP-ribosyl)ation develop normally but are
susceptible to skin disease. Genes & Development, 9(5), 509–520.
Weidele, K., Beneke, S., & B€ urkle, A. (2017). The NAD(+) precursor nicotinic acid
improves genomic integrity in human peripheral blood mononuclear cells after
X-irradiation. DNA Repair, 52, 12–23. https://doi.org/10.1016/j.dnarep.2017.02.001.
Weidele, K., Kunzmann, A., Schmitz, M., Beneke, S., & B€ urkle, A. (2010). Ex vivo supple-
mentation with nicotinic acid enhances cellular poly(ADP-ribosyl)ation and improves
cell viability in human peripheral blood mononuclear cells. Biochemical Pharmacology,
80(7), 1103–1112. https://doi.org/10.1016/j.bcp.2010.06.010.
Williams, P. A., Harder, J. M., Foxworth, N. E., Cochran, K. E., Philip, V. M., Porciatti, V.,
et al. (2017). Vitamin B3 modulates mitochondrial vulnerability and prevents glaucoma
in aged mice. Science (New York, NY), 355(6326), 756–760. https://doi.org/10.1126/
science.aal0092.
Winstall, E., Affar, E. B., Shah, R., Bourassa, S., Scovassi, I. A., & Poirier, G. G. (1999).
Preferential perinuclear localization of poly(ADP-ribose) glycohydrolase. Experimental
Cell Research, 251(2), 372–378. https://doi.org/10.1006/excr.1999.4594.
Wu, Z.-H., & Zhao, S.-P. (2009). Niacin promotes cholesterol efflux through stimulation of
the PPARgamma-LXRalpha-ABCA1 pathway in 3T3-L1 adipocytes. Pharmacology,
84(5), 282–287. https://doi.org/10.1159/000242999.
Yahata, N., Yuasa, S., & Araki, T. (2009). Nicotinamide mononucleotide adenylyltransferase
expression in mitochondrial matrix delays Wallerian degeneration. The Journal of Neuro-
science: The Official Journal of the Society for Neuroscience, 29(19), 6276–6284. https://doi.
org/10.1523/JNEUROSCI.4304-08.2009.
Yamazaki, F., Kuroiwa, T., Takikawa, O., & Kido, R. (1985). Human indolylamine 2,3-
dioxygenase. Its tissue distribution, and characterization of the placental enzyme. The
Biochemical Journal, 230(3), 635–638.
Yamboliev, I. A., Smyth, L. M., Durnin, L., Dai, Y., & Mutafova-Yambolieva, V. N. (2009).
Storage and secretion of beta-NAD, ATP and dopamine in NGF-differentiated rat pheo-
chromocytoma PC12 cells. The European Journal of Neuroscience, 30(5), 756–768. https://
doi.org/10.1111/j.1460-9568.2009.06869.x.
Yang, C.-S., Jividen, K., Spencer, A., Dworak, N., Ni, L., Oostdyk, L. T., et al. (2017).
Ubiquitin modification by the E3 ligase/ADP-ribosyltransferase Dtx3L/Parp9. Molecular
Cell, 66(4), 503–516.e5. https://doi.org/10.1016/j.molcel.2017.04.028.
Yang, H., Yang, T., Baur, J. A., Perez, E., Matsui, T., Carmona, J. J., et al. (2007). Nutrient-
sensitive mitochondrial NAD+ levels dictate cell survival. Cell, 130(6), 1095–1107.
https://doi.org/10.1016/j.cell.2007.07.035.
Ying, W. (2008). NAD +/NADH and NADP+/NADPH in cellular functions and cell
death: Regulation and biological consequences. Antioxidants & Redox Signaling, 10(2),
179–206. https://doi.org/10.1089/ars.2007.1672.
Yu, S. W., Andrabi, S. A., Wang, H., Kim, N. S., Poirier, G. G., Dawson, T. M., et al.
(2006). Apoptosis-inducing factor mediates poly(ADP-ribose) (PAR) polymer-induced
cell death. Proceedings of the National Academy of Sciences of the United States of America,
103(48), 18314–18319. https://doi.org/10.1073/pnas.0606528103.
Niacin 149

Zeman, M., Vecka, M., Perlı́k, F., Staňková, B., Hromádka, R., Tvrzická, E., et al. (2016).
Pleiotropic effects of niacin: Current possibilities for its clinical use. Acta Pharmaceutica
(Zagreb, Croatia), 66(4), 449–469. https://doi.org/10.1515/acph-2016-0043.
Zhang, X., Kurnasov, O. V., Karthikeyan, S., Grishin, N. V., Osterman, A. L., & Zhang, H.
(2003). Structural characterization of a human cytosolic NMN/NaMN
adenylyltransferase and implication in human NAD biosynthesis. The Journal of Biological
Chemistry, 278(15), 13503–13511. https://doi.org/10.1074/jbc.M300073200.

You might also like