Next Article in Journal
Effects of P:Ni Ratio on Methanol Steam Reforming on Nickel Phosphide Catalysts
Next Article in Special Issue
5-((3-Amidobenzyl)oxy)nicotinamides as SIRT2 Inhibitors: A Study of Constrained Analogs
Previous Article in Journal
Antimicrobial Potential of Aqueous Extract of Giant Sword Fern and Ultra-High-Performance Liquid Chromatography–High-Resolution Mass Spectrometry Analysis
Previous Article in Special Issue
A Versatile Continuous Fluorometric Enzymatic Assay for Targeting Nicotinate Phosphoribosyltransferase
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nicotinamide Riboside, a Promising Vitamin B3 Derivative for Healthy Aging and Longevity: Current Research and Perspectives

by
Andrei Biţă
1,2,
Ion Romulus Scorei
2,*,
Maria Viorica Ciocîlteu
2,3,
Oana Elena Nicolaescu
4,
Andreea Silvia Pîrvu
5,
Ludovic Everard Bejenaru
1,
Gabriela Rău
2,6,
Cornelia Bejenaru
7,
Antonia Radu
7,
Johny Neamţu
2,8,
George Dan Mogoşanu
1,2 and
Steven A. Benner
9
1
Department of Pharmacognosy & Phytotherapy, Faculty of Pharmacy, University of Medicine and Pharmacy of Craiova, 2 Petru Rareş Street, 200349 Craiova, Dolj County, Romania
2
Department of Biochemistry, BioBoron Research Institute, S.C. Natural Research S.R.L., 31B Dunării Street, 207465 Podari, Dolj County, Romania
3
Department of Analytical Chemistry, Faculty of Pharmacy, University of Medicine and Pharmacy of Craiova, 2 Petru Rareş Street, 200349 Craiova, Dolj County, Romania
4
Department of Pharmaceutical Technology, Faculty of Pharmacy, University of Medicine and Pharmacy of Craiova, 2 Petru Rareş Street, 200349 Craiova, Dolj County, Romania
5
Department of Biochemistry, Faculty of Medicine, University of Medicine and Pharmacy of Craiova, 2 Petru Rareş Street, 200349 Craiova, Dolj County, Romania
6
Department of Organic Chemistry, Faculty of Pharmacy, University of Medicine and Pharmacy of Craiova, 2 Petru Rareş Street, 200349 Craiova, Dolj County, Romania
7
Department of Pharmaceutical Botany, Faculty of Pharmacy, University of Medicine and Pharmacy of Craiova, 2 Petru Rareş Street, 200349 Craiova, Dolj County, Romania
8
Department of Physics, Faculty of Pharmacy, University of Medicine and Pharmacy of Craiova, 2 Petru Rareş Street, 200349 Craiova, Dolj County, Romania
9
Foundation for Applied Molecular Evolution (FfAME), 13709 Progress Avenue, Room N112, Alachua, FL 32615, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(16), 6078; https://doi.org/10.3390/molecules28166078
Submission received: 27 June 2023 / Revised: 9 August 2023 / Accepted: 14 August 2023 / Published: 15 August 2023
(This article belongs to the Special Issue NAD Metabolism and Related Drug Compounds)

Abstract

:
Many studies have suggested that the oxidized form of nicotinamide adenine dinucleotide (NAD+) is involved in an extensive spectrum of human pathologies, including neurodegenerative disorders, cardiomyopathy, obesity, and diabetes. Further, healthy aging and longevity appear to be closely related to NAD+ and its related metabolites, including nicotinamide riboside (NR) and nicotinamide mononucleotide (NMN). As a dietary supplement, NR appears to be well tolerated, having better pharmacodynamics and greater potency. Unfortunately, NR is a reactive molecule, often unstable during its manufacturing, transport, and storage. Recently, work related to prebiotic chemistry discovered that NR borate is considerably more stable than NR itself. However, immediately upon consumption, the borate dissociates from the NR borate and is lost in the body through dilution and binding to other species, notably carbohydrates such as fructose and glucose. The NR left behind is expected to behave pharmacologically in ways identical to NR itself. This review provides a comprehensive summary (through Q1 of 2023) of the literature that makes the case for the consumption of NR as a dietary supplement. It then summarizes the challenges of delivering quality NR to consumers using standard synthesis, manufacture, shipping, and storage approaches. It concludes by outlining the advantages of NR borate in these processes.

Graphical Abstract

1. Introduction

Nicotinamide riboside (NR) is widely used as a dietary supplement. Structurally, it is a form of vitamin B3 (nicotinic acid, niacin, NA), incorporating into its structure more elements of nicotinamide adenine dinucleotide (in its oxidized form, NAD+) [1]. NR influences, in particular, energy metabolism and neuroprotection [2,3,4].
From a non-medical perspective, as a part of the NAD+ cofactor, NR is also inferred to be a vestige of the “ribonucleic acid (RNA) world” [5], an episode of life on early Earth where RNA was the only encoded component of both metabolism and genetics. This inference is supported by the ubiquity of NAD+ in all branches of the tree of modern life [6].
NR is, however, a rather reactive molecule. Its glycosidic bond joins a positively charged pyridinium heterocycle to a carbohydrate. This bond is therefore especially unstable to cleavage, making NR difficult to synthesize, store, and transport. This creates broad utility for any method to synthesize NR, as well as broad utility for derivatives of NR that are more stable.
Stable forms of NR are especially important today. In July 2013, NR became accessible in dietary supplements in the form of NR chloride (NRCl). This is sold and widely consumed as a dietary supplement under various trademarks, including Tru Niagen™ and Niagen® (produced by ChromaDex, Los Angeles, CA, USA). Other products containing NRCl have been commercially available [7,8]. Interestingly, analysis of some of these products at various times has shown that they do not contain pure NR but also products undoubtedly arising from the reactivity of NR.
Notwithstanding, commercial NR products are suggested to elevate the level of NAD+ in those who consume it [8]. Studies report that chronic NR supplementation (NRS) is well tolerated and elevates NAD+ in healthy middle-aged and older adults [2,3]. Other studies suggest NR oral bioavailability in mice and humans [9]. NR as a dietary supplement has also been the target of clinical studies [2,3]. These have proven its effect of boosting NAD+ in its consumers [8,10]. These facts make it timely to review the current state of the art.
NR, a pyridine-nucleoside form of vitamin B3, consists of nicotinamide (NAM) and ribose as its fundamental components. It is found in various sources, such as milk, yeast, beer, bacteria, and mammals. NR-enriched foodstuffs per se are not well established. Presumably, products that contain yeast are excellent natural sources for the compound [11,12]. However, dairy products have also been noted to contain NR [9,13]. In general, the amounts of NR in foodstuffs are quite low, most likely at micromolar levels.
NR shares biological properties with other NAD+ precursors, specifically NAM and NA. Indeed, these are old or “classical” vitamins B3. Unlike NR, NAM and NA have disadvantages when consumed. Hepatotoxicity can be side effect associated with NAM, whereas a recent preclinical study indicates that NAM remains in the rat body for a shorter duration when compared to NR. NA taken in large amounts is associated with adverse effects, including cutaneous flushing when administered in an immediate release formulation. Sustained release formulations may cause hepatotoxicity.
A precursor that is structurally more advanced toward NAD+ is nicotinamide mononucleotide (NMN). NMN has been recently discovered to be converted extracellularly to NR, which is transported into cells [11,14,15]. Since NMN is converted to NR in the body and the price of NMN is half of the price of NR, technology producing NR to replace NMN as a dietary supplement is especially needed [16]. Among the NAD+ precursors, NR may be preferred, as it produces fewer reported unfavorable side effects [7].
In its classical metabolic roles, NAD+ is a versatile receiver of hydride equivalents, becoming NADH. Analogous chemistry is seen with its phosphorylated derivative, nicotinamide adenine dinucleotide phosphate (NADP+ to NADPH). Traditionally, NAD+ and its derivatives function as coenzymes for dehydrogenases and oxidoreductases, playing crucial roles in fundamental energy metabolism processes, such as glycolysis, the citric acid cycle, and mitochondrial (MT) electron transport. In addition to its classical functions, NAD+ also serves as a vital substrate for signaling enzymes, e.g., poly [adenosine diphosphate (ADP)-ribosyl] polymerases, sirtuins (SIRTs), and ADP-ribosyl-transferases. These enzymes are referred to as “NAD+ consumers” [17,18,19,20,21].
As is classically understood, as well as from recent discoveries, NAD+ is seen as a crucial and abundant metabolite present in all mammalian cells. It participates in a wide array of cellular mechanisms, including essential processes such as metabolism and cell signaling that are vital for survival. This notwithstanding, the limited presence of NR in food sources (with some quantitative studies available), together with the challenges associated with obtaining significant quantities of pure NR as a dietary supplement, have hindered research on the effects of NR on cells and tissues [22,23]. Nevertheless, recently, synthesis techniques for producing NR have been notably advanced, resulting in new and refined approaches [24]. However, these new technologies were developed for cell-based investigations and animal feeding tests, not for human consumption [25,26].
Non-classical roles also exist for NAD+. For example, NAD+ activates SIRTs and supports the MT response to unfolded proteins. Pursuing these non-classical roles, NAD+ metabolism is seen to be involved in an extensive spectrum of pathologies, including cancer, neurodegenerative disorders, cardiomyopathy, obesity, diabetes, and even hearing loss [4,27,28,29,30,31].
In many tissues (e.g., brain, muscle, skin, liver, pancreas, and adipose tissue), the level of NAD+ decreases with age. A separate decrease in the NAD+/NADH ratio affects the cellular redox state, highlighted by anaerobic glycolysis and oxidative phosphorylation (OXPHOS). This decreases, it is thought, the capacity of cells to generate adenosine triphosphate (ATP) [32]. Consequently, healthy aging and longevity are thought to be related to NAD+ metabolism, mainly through NR and NMN, two essential and well-studied NAD+ derivatives. One argument for NAM-related dietary supplements is that these can have prophylactic and therapeutic impact on functional decline, improving age-associated neurodegenerative, cardiovascular (CV), and metabolic diseases and conditions, and promoting the beneficial effects of calorie restriction (CR) [4,33,34,35,36].
As a precursor of NAD+, NR is also proposed to be important in regulating oxidative stress (OXS), inhibiting oxidative injury and inflammatory response, with beneficial effects in the treatment of sepsis [37], systemic lupus erythematosus (SLE) [38], and necrotizing enterocolitis (NEC) [39]; furthermore, NR showed significant protective effect on lung injury caused by paraquat (PQ) in mice [40].
Separately, sepsis-caused multiple organ failure is the major risk of morbidity and mortality in intensive care units. In experimental models of mouse sepsis induced by lipopolysaccharide (LPS) injection, feces injection in peritoneum, or by cecal ligation and puncture, NR inhibited plasma high mobility group box 1 (HMGB1) release, OXS, and tissue infiltration, increased endogenous antioxidant ability, prevented lung and heart injury, and improved survival [37,41]. This is mediated via NAD+/SIRT1 signaling.
NR could be a potential adjuvant for SLE treatment. In vitro testing in monocytes from patients with SLE showed that NR restricted autophagy (ATG) and attenuated interferon-beta (IFN-β) release in an NAD+-dependent manner, but also through inosine signaling [38]. Also, in the mouse experimental NEC model, NR administration alleviated OXS, increased NAD+ levels and intestinal microcirculatory perfusion, and relieved signs of endothelial dysfunction by modulating the SIRT1-associated endothelial nitric oxide synthase (eNOS) acetylation/deacetylation pathway [39,42].
Compared with the control (CON) group, which received saline by one-time gavage, intraperitoneal (i.p.) injection of 300 mg/kg NR solution led to the inhibition of the inflammatory response, peroxidation injury, and apoptosis at the lung level and to the survival time prolongation of the PQ intoxicated mice, mainly by upregulating SIRT1 and nuclear factor erythroid 2 (NF-E2)-related factor 2 (NRF2) protein expression [40].
We recently reported an inexpensive synthesis of borate-stabilized NR based on models for how nucleosides might have been formed on a prebiotic Earth. Unlike many commercially available NR products, borate-stabilized NR can be delivered in pure form, where it exhibits high stability against thermal and chemical decomposition. Thus, NR borate is a promising nutraceutical to replace NRCl, as currently sold on the United States (U.S.) market. This paper reviews the current research supporting the value of NR for healthy aging and longevity, the perspectives of senotherapeutic NAD+ supplementation, and thus forms of NR that can be made, shipped, and stored without decomposition.

2. Effects of Nicotinamide Riboside on Several Organs and Systems

2.1. Nicotinamide Riboside and the Nervous System

The effects of NR on energy metabolism and neuroprotection were highlighted with the first research about this compound. In yeast and mammals, NR is metabolized by two main pathways: (i) degradative processes from which NAM results; (ii) synthetic processes exploiting NR kinases (NRK1, NRK2), resulting in increased tissue NAD+. This increase is reported to increase insulin sensitivity, MT biogenesis, and SIRT functions.
Thus, in experimental models of Alzheimer’s disease (AD), NR in larger amounts than is found naturally in food leads to brain-protective effects by the stimulation of NAD+ anabolism [7]. NR is a neuroprotective factor that improves cognition after brain injuries, such as acute ischemia. In an experimental model, 300 mg/kg NRCl was i.p. administered, 20 min after reperfusion, in mice with middle cerebral artery (MCA) occlusion. The NRCl-treated group exhibited better memory function and recovery of learning in the Morris water maze test.
Following acute treatment with NRCl, apoptosis in the hippocampus, neuronal loss, and hippocampal infarct volume were decreased. NRCl also increased the measured amounts of NAD+ and ATP, and stimulated adenosine 5′-monophosphate (AMP)-activated protein kinase and SIRT1, as observed by high-performance liquid chromatography (HPLC) and Western blot assays [43]. Moreover, the clinical phenotype and T-cell survival/function were improved by S-adenosylmethionine (SAM) and NR co-administration for patients with phosphoribosyl pyrophosphate synthetase 1 (PRPS1) deficiency (Arts syndrome, manifested by serious neurological and immunological deficiencies especially in males) [44].
NR is reported to be beneficial in other neurological syndromes. For example, Gulf War illness (GWI) is a chronic neuropsychiatric disorder characterized by neurocognitive deficits arising from OXS, neuroinflammation, and neuronal damage. Currently, no effective treatment of GWI is known [45]. However, in a GWI experimental model, NR-mediated SIRT1 activation restored brain MT bioenergetics dysfunction (following astroglia activation) and reduced neuroinflammation. Here, NR was administered as a dietary supplement, 100 μg/kg daily for two months. In the brain of NR-treated GWI mice, a decrease in lipid peroxidation and proinflammatory cytokines was closely correlated to the increase in deacetylation of the peroxisome proliferator-activated receptor gamma coactivator-1 alpha (PGC-1α) and nuclear factor-kappaB (NF-κB) p65 subunit [46].
Many central nervous system (CNS) disorders arise from dysfunctionalities in the gut microbiota–brain axis. In the mice brain, NRS provided protection against alcohol-induced depressive behavior and decreased the level of anti-inflammatory (interleukin (IL)-10 and transforming growth factor (TGF)-β) and proinflammatory (IL-1β, IL-6, and tumor necrosis factor (TNF)-α) cytokines. In the hippocampus, NR significantly reduced the number of activated microglia and the inhibition of the protein kinase B (AKT)/glycogen synthase kinase 3 beta (GSK3β)/β-catenin signaling pathway and increased the brain-derived neurotrophic factor (BDNF). Similar with donor mice, in recipient mice, fecal microbiota transplantation (FMT) improved the microglial activity, the level of BDNF and cytokines, the activation of AKT/GSK3β/β-catenin signaling pathway, and cognitive behavior [47].
Intracortical administration of NR protects against excitotoxicity-induced axonal degeneration (AxD) and decreases the brain damage provoked by injection of N-methyl-D-aspartate (NMDA). Both NR and NAD+ prevented neuronal death due to the axonal stress. However, NR exhibited better neuroprotection than NAD+ at the level of cortical neurons [48].
Moreover, in rats with TNF-induced optic nerve degeneration, intravitreal injection of NR showed significant axonal protection. In retina and optic nerve, NR upregulated the levels of SIRT1 ATG pathway, decreased the p62 protein levels and increased the levels of microtubule-associated protein 1A/1B-light chain 3-II (LC3-II) that localizes with the MT inner membrane. The presence of NRK1, upregulated by NR administration, was also evidenced in the retinal and optic nerve fibers and in the retinal ganglion cells (RGCs) [49,50].
This research continues. For example, Sun et al. (2023) recently showed that NAD+ depletion mediates cytotoxicity in human neurons with ATG deficiency, leading to cytotoxicity and neurodegeneration. The research was conducted on ATG-deficient (ATG5−/−) human embryonic stem cells (hESCs) representing a “human neuronal platform”. In ATG5−/− neurons, NR treatment improved cell viability by the restoration of MT bioenergetics and proteostasis [51].
In experimental (preclinical) models and also in clinical studies, NR improved ataxia scores and immunoglobulin G (IgG) levels in ataxia telangiectasia (A–T). In ATM-deficient mice, which is a model for the A–T phenotype in humans, NR administration improved motor function and prevented neuroinflammation, neurodegeneration, MT dysfunction and cellular senescence [52]. NR dietary supplementation was well tolerated by A–T patients; ataxia, dysarthria, and the quality of life (QoL) were all reportedly improved without adverse effects [53,54].
In cluster of differentiation (CD) 157 knockout (KO) male mice, oral administration of NR increased the levels of oxytocin (OT) in cerebrospinal fluid (CSF), stimulating the release of this anxiolytic factor during stress, and corrected the fearful and anxiety-like behaviors and the social deficits [55]. A recent study highlighted the remission of social behavior impairment by oral gavage administration of NR for 12 days in CD157, but not in CD38 KO mice [56].
In cell culture (in vitro) and in mouse models of amyotrophic lateral sclerosis (ALS), NR and pterostilbene (PT) supplementation cooperatively delayed motor neuron failure, decreased the levels of neuroinflammation markers in the spinal cord, influenced the muscle metabolism, and increased to a small extent the survival of hSOD1G93A transgenic mice [57]. Further, in SOD1G93A transgenic and wild-type (WT) mice, starting at 50 days of age, oral administration of NR (20 mg/mL) improved MT proteostasis and adult neurogenesis. In the brain of SOD1G93A mice, neural stem cells (NSCs)/neuronal precursor cells (NPCs) proliferation and migration were enhanced following the NRS [58]. A recent clinical trial evidenced that NR and PT co-administration proved to be effective in inhibiting OXS-induced cellular damage, the major pathophysiological mechanism of ALS [59,60].
In mouse models for Cockayne syndrome (CS) (CSA−/− and CSBm/m, modeling human CS), NR is reported to protect from noise-induced hearing loss (NIHL). The main mechanism of action involves the activation of NAD+/SIRT3 pathway, which contributes to the reduction in spiral ganglia neurite degeneration caused by intensive noise exposure (NE). NAD+ supplementation could prove beneficial for the treatment of CS-related hearing loss at a young age and age-related hearing loss (ARHL) affecting elderly individuals [61,62,63].
In Sprague–Dawley female rats, oral administration of NR (200 mg/kg) for four weeks relieved the nociceptive and aversive dimensions of intravenous (i.v.) paclitaxel (PTX)-induced peripheral neuropathy. In addition, NRS increased NAD+ plasma level by 50% and did not influence the myelosuppressive properties of PTX, the adverse locomotor effects not being observed [64].
In male rats, NR oral treatment (500 mg/kg) alleviated PTX-induced corneal and somatic hypersensitivity to tactile stimuli, with no suppression of basal tear production/chemosensitivity and without altering the corneal afferent density. NR also reversed the PTX-induced hindpaw hypersensitivity to cool and tactile stimuli, with no inverting of the non-peptidergic intraepidermal nerve fiber (IENF) loss [65].
In the recent studies regarding neuroprotective therapy for glaucoma, in a mouse model of acute RGC damage caused by optic nerve crush (ONC), and with chronic RGC degeneration obtained by intracameral injection of microbeads that induced ocular hypertension, orally administered NR removed retinal inflammation, as shown by immunofluorescence staining for glial fibrillary acidic protein (GFAP). It also enhanced survival and preserved the function of RGC [66,67]. In a mouse model of light-induced retinal degeneration, i.p. injected NR increased the NAD+ level and had a protective effect on retinal function through the restoration of the photoreceptor cell layers, by the reduction in the inflammation and by the limitation of the consequences of apoptosis process [68].
In 125 patients with primary open-angle glaucoma (POAG), a randomized double-blind, placebo (PLA)-controlled trial showed an effect of NR (300 mg for 24 months) in slowing down optic nerve degeneration. Compared with PLA, NR patients had a lower degree of progressive retinal nerve fiber layer (RNFL) thinning and visual field (VF) loss [69,70].
The protective effects of NR were also investigated on hydrogen peroxide-induced oxidative damage in lens epithelial cells. In vitro, on SRA01/04 cell line, NR significantly reduced apoptosis and the generation of reactive oxygen species (ROS), increased cell viability, and improved levels of superoxide dismutase (SOD), catalase (CAT), total glutathione (GSH), and MT membrane potential. NR was found to reduce OXS damage through the targeting of mitogen-activated protein kinase (MAPK) and Janus kinase 2 (JAK2)/signal transducer and activator of transcription 3 (STAT3) pathways [71].
Pang et al. (2021) reported that some aberrant metabolic pathways during pregnancy, such as NAD+, OXPHOS, tricarboxylic acid (TCA) cycle and tryptophan metabolism, underlie Zika virus (ZIKV)-induced microcephaly in newborns. Preclinical experiments suggested that in ZIKV-infected mice, NRS improved survival, reduced the apoptosis, and increased the thickness of the cerebral cortex [72].

2.2. Nicotinamide Riboside and the Cardiovascular System

Dietary supplementation by NAD+ precursors (e.g., NR) appears to have a useful impact in many areas of cardiovascular diseases (CVDs). Here, the goal is to improve overall cardiometabolic health by increasing OXPHOS capacity and mitophagy, removing mitochondrial deoxyribonucleic acid (mtDNA) mutations, and for cardioprotective effects in ischemia–reperfusion (I–R) injury, arrhythmias, heart failure (HF), myocardial infarction (MI), and high blood pressure [73,74,75,76].
For example, in an experimental model of cafeteria (CAF) diet-induced obesity (DIO), oral administration of 400 mg/kg of NR combined with CR (−62% kcal) for 28 days decreased the weight of obese male Wistar rats, their visceral and subcutaneous adiposity, their triglyceride (TG)/high-density lipoprotein (HDL) ratio, and their heart size. Also, insulin resistance and antioxidant capacity (glutathione peroxidase (GPx) and CAT enzymatic markers of cardiac OXS) were improved by NRS [77]. In a mouse model, NR treatment for 30 days led to the NAD+/NADH ratio balancing, lowering acetylation level, improving MT function and HF with preserved ejection fraction (EF) phenotypes [78]. In a murine model of myocardial hypertrophy induced by transverse aortic constriction (TAC) surgery, NR dietary supplementation reduced levels of inflammatory cytokines (IL-1β, TNF-α), mitigated the NLR family pyrin domain containing 3 (NLRP3) inflammasome activation, the elevation of myocardial NAD+ level, and the improving of cardiac dysfunctions and morpho-functional changes (myocardial hypertrophy). The research led to the observation that the NAD+/SIRT3/manganese superoxide dismutase (MnSOD) signaling pathway was also regulated by NRS [79]. Moreover, in mice, NR administration inhibited TAC-induced endothelial-to-mesenchymal transition of endothelial cells (ECs) and promoted MT unfolded protein response leading to improvement in prohibiting proteins’ expression and to the decrease in cardiac fibrosis (CF) progression [80].
Recent clinical trials highlighted the cardioprotective role of oral NRS (usually 1 g twice daily, a well-tolerated dose) for ischemic heart disease patients diagnosed with atrial fibrillation (AF) [81], HF patients with MT dysfunction and peripheral blood mononuclear cells (PBMCs) inflammatory activation [82], or clinically stable HF patients with reduced EF [83].
In murine models, boosting NAD+ levels using dietary NRS appeared to reduce the development of aortic aneurysms and sudden death by aortic ruptures. Acute aortic aneurysms and lethal ruptures were induced by angiotensin II (Ang II) administration to apolipoprotein E (ApoE)-deficient mice fed throughout the entire experiment with a Western diet (WD). Fatal aortic ruptures caused by atherosclerosis and hypertension are reported to be prevented by boosting MT respiration [84].
Additional studies in this space have accumulated. Thus, NRK2 is reported to limit dilated cardiomyopathy (DCM) in mice with chronic pressure overload (PO). In addition, Ang II-induced cardiomyocyte death is reported to be mitigated by NRK2 overexpression [85]. NR administration preserves cardiac function in a mouse model of DCM and TAC-induced cardiac hypertrophy, alleviating HF development by stabilizing NAD+ levels in the myocardial tissue.
NA adenine dinucleotide, methylnicotinamide (MeNAM), and N1-methyl-4-pyridone-5-carboxamide represent the three biomarkers with increased myocardial levels due to NRS [86]. In murine models, NR was i.v. administered to assess its ability to mitigate doxorubicin (DOX)-induced cardiomyopathy. OXS, autolysosomes accumulation and autophagic flux blockade because of DOX cardiotoxicity were all prevented by NR via NAD+/SIRT1 signaling pathway [87,88]. For the specific pathogen-free male Wistar rats, the protective effect exhibited significant levels through the preventive use of NR [89].
NR also attenuated cardiac and I–R injury. In the ischemic heart, NRK2 regulates metabolic adaptation and MT function post-MI [90]. In a cardiac I–R injury experimental (mouse) model, orally administered NR and resveratrol (RSV) nanocrystal self-assembled microspheres (NR/RSVms) for eight hours decreased MI, with no important adverse effects on internal organs [91]. Also, in a fentanyl–midazolam anesthetized I–R injury rat model, NR was i.v. administered as a bolus before the ischemia started. It was found that NR could target cardiac I–R injury as a promising cardioprotective natural compound [92]. Moreover, in a mouse model of superior mesenteric artery ischemia, NRS improved microcirculation and mesenteric vessels relaxation and protected the intestinal wall against I–R injury [93].
In human aortic endothelial cells (HAECs) and in murine aortic rings, NR acts as a vasoprotective agent at the endothelial level by modulation of intracellular NAD+ and inhibition of inflammation, as evidenced by the decreased expression of von Willebrand factor (vWF) and intercellular adhesion molecule-1 (ICAM-1) [94].

2.3. Nicotinamide Riboside and the Digestive System

In mice subjected to partial hepatectomy, boosting NAD+ by putting NRS in the drinking water promoted liver regeneration. This was interpreted as increasing DNA biosynthesis at the hepatocyte level and amelioration of hepatic steatosis (HS) [95]. In a high-fat diet (HFD)-fed murine model, NR treatment enhanced hepatic MT function and NAD+ levels and prevented lipid accumulation in the liver [96]. In vitro, HS was induced in AML12 mouse hepatocytes treated with 250 μM palmitic acid for 48 h. Cell exposure to NR (10 μM and 10 mM) for 24 h did not affect morphology nor viability; instead, it was observed a decrease in TNF-α and IL-6 proinflammatory markers and an increase in SIRT1 gene activity, PGC-1α, transcription factor A, carnitine palmitoyltransferase 1, uncoupling protein 2, mtDNA and MT biogenesis [97].
Supplementation with NAD+ precursors (NR, NMN) can prevent liver injury. Proposals to treat non-alcoholic fatty liver disease (NAFLD) NRK1 by targeting HS are based on the discovery that in aged mice or in mice treated with HFD, NRK1 overexpressed by adenovirus-mediated gene transduction regulates lipid metabolism (mainly TG level) and NAD+ biosynthesis in the liver, insulin sensitivity, and glucose tolerance [98]. In aged mice with a moderate NAFLD-like phenotype, NRS (2.5 g/kg food for three months) significantly reduced the levels of total cholesterol (TC), TG, aspartate aminotransferase (AST), and alanine aminotransferase (ALT) and increased the amount of liver NAD+, diminished inflammatory infiltration, and mitigated HS and liver fibrosis (LF) [99,100,101]. In a diet-induced NAFLD mice model, supplementation with a multi-ingredient mixture of betaine, N-acetylcysteine, L-carnitine, and NR slowed NAFLD progression and influenced the gut–liver axis through the correction of intestinal microbiota dysbiosis and modulation of short-chain fatty acids (SCFAs) levels in feces [102].
Daily NR and PT (NRPT, Basis™) co-administration in 111 adults with NAFLD was assessed in a six-month prospective, randomized, double-blind, PLA-controlled clinical trial. Here, NRPT treatment was well tolerated, and significantly decreased serum levels of ceramide 14:0 (a toxic lipid), ALT, and gamma-glutamyltransferase (GGT) [103].
In an HFD-induced male C57BL/6J mice model of LF, NR administration (400 mg/kg/day for 20 weeks) attenuated the development of LF. The body weight, the amount of collagen in the liver, and the activation of hepatic stellate cells (HepSCs) were significantly reduced by NRS, compared to liver inflammation, HS, and ALT levels, which have not been mitigated [104]. In a similar context, NR exhibited a protective effect against LF induced by carbon tetrachloride in mice through an increase in SIRT1 gene activity, modulation of Smads signaling pathway acetylation, and suppression of TGF-β-induced HepSCs activation [105].
NR treatment (400 mg/kg for 16 days) attenuated liver injuries induced by alcohol in C57BL/6J mice fed a Lieber–DeCarli ethanol liquid diet. The protective effect of NRS involves SIRT1 activation, increasing hepatic NAD+ levels and PGC-1α deacetylation, enhancing MT biogenesis/functionality, and decreasing OXS status and ethanol-induced lipid accumulation [106].
A recent in vitro study in a human hepatic (HepG2) cell line model of citrin deficiency highlighted the potential use of NRS for balancing dysregulated glycolysis and fatty acid β-oxidation [107].
NRS has even been suggested as being useful to manage the sequelae of infectious diseases. For example, as part of research aimed at new treatments for severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) disease, it was highlighted that NRS significantly inhibited the murine hepatitis virus (MHV) replication by increasing the level of NAD+, activation of the TCA cycle and of MT metabolism [108].

2.4. Nicotinamide Riboside and the Urinary System

In the U.S. hospitals, about 3–10% of the adults suffer from acute kidney injury (AKI). NAD+ supplementation, by oral administration of NAM, NA, NR, dihydronicotinamide riboside (reduced NR, NRH) and NMN, and SIRT activation may be useful in the prophylaxis and treatment of AKI, if some recent studies on MT metabolism are developed [109,110,111,112,113]. For example, in an experimental model of newborn (one day) mice from mothers fed a low protein diet, NRS restored the SIRT3 expression, by the induction of PGC-1α, improved MT and cellular protection against OXS, increasing the density of nephrons, renal capillaries, and glomerular podocytes [114].
The effect of four NRPT doses (NR/PT—1.25 g/50 mg, 2.5 g/100 mg, 3.75 g/150 mg, 4 g/200 mg), administered twice a day for two days, was tested in a randomized, double-blind, PLA-controlled study in 24 hospitalized AKI patients. Serum NAD+ levels were increased by NRPT treatment. Safety was assessed by various assays, such as estimated glomerular filtration rate (eGFR), creatinine levels, electrolyte levels, hepatic function parameters, and blood count; these all remained unchanged. Minor gastrointestinal (GI) side effects were reported only by three patients. NRPT supplementation up to 1 g/200 mg twice a day was reported as being safe and well tolerated in hospitalized AKI patients [110].
In another recent preclinical study, in rats, NR prophylactic supplementation did not mitigate tubular impairment and induction of profibrotic genes neither in bilateral I–R injury-induced long-term AKI nor in chronic kidney disease (CKD) experimental model, even though an increased NAD+ blood levels were evidenced [115]. However, Doke et al. (2023), in a study in male mice with cisplatin (CPT)- or I–R-induced AKI, suggested that NR and NMN supplementation restored NAD+ serum levels and improved MT metabolism and kidney functionality [116].
A randomized PLA-controlled, double-blind, crossover trial was conducted for six weeks to determine the impact of NR (1 g/day) and coenzyme Q10 (CoQ10, 1.2 g/day) on exercise tolerance and metabolic profile in 25 CKD patients with eGFR of <60 mL/min/1.73 m2. Compared with PLA, only systemic MT metabolism and lipid profile (e.g., TG, free fatty acids, ceramides) were improved by NR and CoQ10 co-administration, but not aerobic capacity (VO2 peak) or total work efficiency [117].

2.5. Nicotinamide Riboside and the Musculoskeletal System

Numerous preclinical studies have suggested that NR and NMN dietary supplements increase NAD+ levels in the skeletal muscle and help protect musculoskeletal system from age-related metabolic dysfunction. NAD+ biosynthesis pathways/homeostasis and its bioavailability in skeletal muscle cells are influenced by NAM phosphoribosyltransferase (NAMPT), NRK1, and NRK2 enzymes [118,119]. After the application of NRS in hexose-6-phosphate dehydrogenase KO (H6PDKO) mice, NAD+/NADH ratio was elevated but without any effect on insulin sensitivity, MT respiratory dysfunctions (MT acylcarnitine), acetyl coenzyme A metabolism and endoplasmic reticulum OXS [120,121].
In a double-blind, PLA-controlled, randomized, crossover trial for 21 days, 12 aged men were supplemented with 1 g/day NR. It was found that both MT bioenergetics and skeletal muscle NAD+ metabolism were stimulated by NRS; at the same time, the amounts of serum inflammatory cytokines were reduced [122]. Moreover, in a randomized, double-blinded, PLA-controlled, crossover intervention study in 13 healthy obese humans, NRS (1 g daily for six weeks) induced some minor modifications in body composition and sleeping metabolic rate, increased the level of NAD+ metabolites, such as NA adenine dinucleotide and MeNAM, and altered the acylcarnitine concentrations in the skeletal muscle [123].
Recent studies showed that, in healthy humans, oral NAD+ supplementation (NR, NMN) does not alter whole-body or skeletal muscle metabolic responses to an endurance exercise [124,125]. During endurance exercises, one week of NRS (1 g daily) for eight young males did influence neither skeletal muscle MT respiration nor auto-PARylation of poly ADP-ribose polymerase 1 (PARP1), acetylation of tumor protein 53 (p53) Lys382 and MnSODLys122, or overall acetylation [126]. A recent randomized controlled trial testing of in-home aerobic and exercise training and 16 weeks NRS targeted the skeletal muscle MT OXPHOS and improved the muscle mass and fitness in adolescent and young adult survivors of hematopoietic cell transplantation (HemSCT) [127].
However, some preclinical studies showed that, compared to the CON group, chronic NRS (300 mg/kg/day for 21 days by oral gavage) decreases the exercise/swimming performance in Wistar rats. The redox features and pleiotropic metabolism of NAD+ and NADP+ could be an explanation for the potential inhibition of oral NR treatment on physical performance of rats [128,129,130].
NAD+ repletion improves MT activity and muscle stem cells (MuSCs) function and enhances life span in aged mice. In an experimental model of muscular dystrophy, NRS prevented the senescence of MuSC in the mdx (C57BL/10ScSn-Dmd(mdx)/J) mice [131,132]. Also, seeking treatment of MT myopathy, orally administered NR was found to induce MT unfolded protein response, MT biogenesis in brown adipose tissue (BAT) and skeletal muscle, prevent abnormalities of MT ultrastructure, and prevent deletion of mtDNA [133]. In this regard, NRS increased the levels of NAD+ in liver and skeletal muscle, exercise capacity, and MT respiration, alleviating the exercise intolerance in adenine nucleotide translocator 1 (ANT1)-deficient mice [134]. In the recent study on 20 body mass index (BMI)-discordant monozygotic twin pairs, NRS (250 mg to 1 g daily for five months) improved muscle MT biogenesis, satellite cell differentiation, and gut microbiota [135].
NAD+ boosting was preclinically characterized as a promising therapeutic approach for rheumatoid arthritis (RA) patients. NR and NMN treatments were ex vivo tested on RA–PBMCs, thus highlighting NAD+ amount increasing via NAMPT and nicotinamide mononucleotide adenylyl transferase (NMNAT), and reducing the proinflammatory, pro-oxidative and pro-apoptosis in RA patients [136].
The influence of NR (250 mM, 500 mM, 1 M) on the development and growth of pectoralis major muscle (PMM) was examined on fertilized Cobb 500 broiler eggs. NRS introduced by injection into the yolk and albumen influenced in ovo broiler myogenesis in a fashion correlated with an increase in PMM weight, length, and fiber density [137,138,139].

3. Nicotinamide Riboside as a Tool to Mitigate Metabolic Disorders

3.1. Nicotinamide Riboside and Obesity

Among NAD+ precursors, NR is often suggested as a dietary supplement to enhance oxidative metabolism and protect against HFD-induced obesity and deficient MT functionality for age-related diseases [25,140]. In aging mice, NAD+ supplementation via NR rejuvenated intestinal stem cells (ISCs) improved the repairing capacity of gut damages by increasing the activity of SIRT1 and mammalian target of rapamycin complex 1 (mTORC1) [141]. In experimental models of mildly obese male mice (C57BL/6N, C57BL/6J), NRS for eight weeks increased MT respiration. However, it had minimal impact on energy metabolism, without influencing body weight, internal organs weight, glucose metabolism, liver lipids amount, and metabolic adaptability [142].
NR- and NMN-conditioned microbiota reduced HFD-induced weight gain in C57BL/6J male mice by increasing their energy consumption. Here, butyrate-producing Firmicutes were enriched by FMT from NR-supplemented donors to HFD-fed naïve mice [143,144]. Alterations in intestinal brush border membrane functionality and bacterial populations following intra-amniotic administration (to Gallus gallus) were reported for 30 mg/mL doses of NR and its derivatives, water-soluble NR tributyrate chloride and oil-soluble NR trioleate chloride. Bifidobacterium, Clostridium, Lactobacillus, and Escherichia coli populations were significantly increased by NR treatment [145]. Further, oral supplementation of NR for 12 weeks alters gut microbial composition (Erysipelotrichaceae, Lachnospiraceae and Ruminococcaceae families) in HFD-fed rats and mice, but not in humans [146].
NRS exerts an anti-obesity effect and prevents inflammation and LF in white adipose tissue (WAT) of 8-week-old (young) and 16-week-old (old) female C57BL/6J DIO mice fed an HFD/high-sucrose diet/high-cholesterol diet or HFD combined with 400 mg/kg/day NR for 20 weeks. Weight and size of gonadal WAT (gWAT) adipocytes of old mice females were decreased by NRS [147]. Similar results were seen in recent research concerning NR and CR effects on gut microbiota and liver inflammatory and morphological markers in CAF DIO in adult male Wistar rats [148]. The effect of NR on lipid metabolism and gut microflora–bile acid axis was studied in alcohol-exposed mice. NRS increased the NAD+/NADH ratio, deoxycholic acid and hyocholic acid levels, and decreased the activation of the protein phosphatase 1 (PP1) signaling pathway, chenodeoxycholic acid, TG and total bile acid levels, and lipid accumulation [149].
Oral NRS also appeared to confer metabolic benefits in obese mice. Thus, NR induces a thermogenic response in the BAT of lean mice. Over five weeks, male C57BL/6J mice were supplemented with NR (400 mg/kg/day), giving a reduction in abdominal visceral fat depots and an increase in body temperature [150,151]. Also, compared with CONs, NRS conferred marginal metabolic benefits, augmenting MT functionality in C57BL/6NJ obese mice without remodeling the skeletal muscle acetyl-proteome [152]. NRS provided daily to suckling male mice improved lipid and energy metabolism in skeletal muscle and liver in adulthood, correlated with an upregulation of SIRT1 and AMP-dependent protein kinase signaling pathways [153]. In obese rats under CR, NRS (400 mg/kg for four weeks) neutralized hypothalamic inflammation by reverting high levels of TNF-α and increases weight loss without altering skeletal muscle mass [154]. Further, high doses (9 g/kg diet) of NRS for 18 weeks induced glucose intolerance and WAT dysfunction in male C57BL/6JRccHsd mice fed a mildly obesogenic diet (MOD) containing low but adequate tryptophan [155,156,157].
In an HFD mouse model, a reduction in obesity and insulin resistance was seen by dual targeting of visceral adipose tissue (VAT) and BAT using a novel combination of metabolic cofactors (NR, N-acetylcysteine, betaine, L-carnitine) orally administered for four weeks [158]. NRS (400 mg/kg daily) ameliorates high-fructose-induced lipid metabolism disorder in C57BL/6J mice via improving fibroblast growth factor 21 (FGF21) resistance in the liver and WAT. NR treatment upregulated the SIRT1–NF-κB pathway, decreasing inflammatory processes and increasing NAD+/NADH ratios [159]. Another study concluded that DNA methylation changes are associated with the programming of WAT browning features by RSV and NR neonatal supplementations in mice [160]. A positive signal was seen with a combined treatment of L-carnitine (0.4%, w/w) and NR (0.3%, w/w). This evidently improved hepatic metabolism and attenuated obesity and HS in HFD-fed Ldlr−/− Leiden mice [161].
As a mechanism to treat metabolic disorders, NR may shift the differentiation of human primary white adipocytes to beige adipocytes. This may impact substrate preference and uncouple respiration through SIRT1 activation and MT-derived reactive species production [162]. In a randomized, double-blinded, PLA-controlled, crossover study in human volunteers (45–65 years, BMI 27–35 kg/m2), NR (1 g daily for six weeks) enhanced in vitro beta-adrenergic BAT activity and did not alter cold-induced thermogenesis [163]. A randomized, PLA-controlled, double-blind, clinical trial in 40 healthy obese (BMI > 30 kg/m2) insulin-resistant men looked at safety, insulin sensitivity, and lipid-mobilizing effects of NR (1000 mg twice daily) dietary supplementation for 12 weeks. It was found that NRS was safe and well tolerated, but did not improve either endogenous glucose generation, oxidation, or elimination, and did not improve insulin sensitivity [2]. Further, similar impacts of NRS were seen on endocrine pancreatic function and incretin hormones in non-diabetic men with obesity [164]. In a follow up study by the same authors, NRS does not alter MT respiration, content, or morphology in skeletal muscle in middle-aged, obese, and insulin-resistant men [165].

3.2. Nicotinamide Riboside and Diabetes

NRS can be used to mitigate type 2 diabetes (T2D) and neuropathy in HFD-fed prediabetic and diabetic male C57BL/6J mice. NR treatment significantly restored liver NADP+ and NADPH amounts and reduced weight gain, blood glucose and HS [166]. NAD+ precursors (NMN, NR) improved MT function in diabetes and prevented experimental diabetic peripheral neuropathy (DPN). The experiments were carried out in streptozotocin (STZ)-induced diabetic rats or mice. NMN was administered for two months (50 or 100 mg/kg i.p. injection on alternate days). HFD-fed mice were supplemented by NR at 150 or 300 mg/kg for two months. NR treatment in HFD-fed mice led to the normalization of adult dorsal root ganglion (DRG) neurons’ functionality [167]. From the point of view of cellular and molecular mechanisms, a recent study in NRK1 KO mice showed that NRK1 protected against HFD and age-induced pancreatic β-cell failure [168]. Also, NRS promoted mitofusin 2 (MFN2)-mediated MT fusion in diabetic hearts through the SIRT1–PGC-1α–peroxisome proliferator-activated receptor alpha (PPARα) pathway [169].
The effects appear to be broad. Supplementation with NR reduces brain inflammation and improves cognitive function in diabetic mice. NR treatment for six weeks of HFD-fed diabetic Institute of Cancer Research (ICR) male mice significantly reduced amyloidogenesis (amyloid beta (Aβ) precursor protein, presenilin 1) and neuroinflammation (IL-1, IL-6, and TNF-α markers) [10]. In STZ-induced diabetic rats, i.p. administered NR every 48 h for six weeks improved enteric neuropathy through myenteric plexus neuroprotection [170].
Other benefits of NRS in experimental murine diabetes refer mainly to the (i) alleviation of hepatic metaflammation by modulating NLRP3 inflammasome in a rodent (8-week-old KK/HlJ male mice) model of T2D [171], (ii) enhancement of endothelial progenitor cell (EPC) function to promote refractory healing of diabetic wounds through mediating the SIRT1–AMP-activated protein kinase (AMPK) pathway [172], and (iii) improvement in fetal growth under hypoglycemia previously induced in pregnant mice [173].

4. Nicotinamide Riboside for Healthy Aging and Longevity

4.1. Nicotinamide Riboside for Healthy Aging

NAD+ metabolism and homeostasis are important in aging and disease [174,175]. Declining NAD+ levels induce the augmentation of hypoxia-inducible factor-1alpha (HIF-1α) and a pseudohypoxic state disrupting PGC-1α/-1β-independent nuclear–MT communication during aging [176].
Thus, in recent preclinical studies, managing NAD+ deficits with NAM, NA, NR, and NMN provided healthy aging [177]. In the Caco-2 cell line and NIAAA mouse model, NAD+ supplementation by NR alleviated intestinal barrier injury induced by ethanol via protecting epithelial MT function [178]. NRS increased the formation of human leukocytes from hCD34+ progenitors in the immunodeficient mice model. Thus, through increased MT clearance, NR potently stimulated the hematopoiesis [179] and the lymphoid potential of Atm−/− and old mice HSCs [180,181]. Moreover, NMN can be used to activate SIRT1 and improve the pathophysiology of diet- and age-induced diabetes in mice [182].
Likewise, reduced NMN is a newly analyzed and potent NAD+ precursor in mammalian cells and mice blood and internal organs [183]. In randomized, double-blind, PLA-controlled clinical trials in healthy middle-aged and older adults, chronic NRS (500 mg twice daily) is safe and well tolerated, and it is recommended for (i) increasing NAD+ level [8] and (ii) treating elevated systolic blood pressure (SBP) and arterial stiffness having initial above-normal (120–159 mmHg) SBP [184].
In aging and disease, a close connection relates NAD+ levels and the activation of SIRTs, as seen in murine models: (i) NR alleviates CPT-induced peripheral neuropathy (CIPN) and neuronal death via SIRT2 activation and enhancement of nucleotide excision repair in Lewis lung carcinoma model [185]; (ii) SIRT3 is required for regeneration of WT or mutant liver but not for the beneficial effect of NR [186]; (iii) SIRT3 deficiency aggravates contrast-induced acute kidney injury (CIAKI) in vitro (HK-2 cells) and in vivo (WT and SIRT3 KO mice) [187]; (iv) NR attenuates inflammation and OXS by activation of SIRT1 and normalization of NAD+/NADH ratio in alcohol-stimulated RAW 264.7 macrophages and in mouse bone marrow-derived macrophages [188].
NRS could be a missing piece in the puzzle of exercise therapy for older adults. Short-term oral NR treatment, 300 mg/kg or 600 mg/kg daily, improves muscle quality and function in middle-aged male C57BL/6J mice and increases cellular energetics and differentiating capacity of myogenic progenitors [189]. A randomized, PLA-controlled trial of safe and well-tolerated daily NRPT supplementation (1 g NR and 0.2 g PT) improved skeletal muscle regeneration after experimental muscle injury in 23 elderly individuals (55–80 years) [190]. Moreover, in a double-blind, crossover study, acute NRS improved redox homeostasis and exercise performance in 12 old individuals compared with the same number of young men [191].
Recent studies have highlighted that senotherapeutic NRS could contribute to healthy aging and longevity. As a sample of results: (i) the effects of senolytic drugs, including NR, were tested on human mesenchymal stromal cells (MSCs) with no significant action on molecular markers for replicative senescence [192]; (ii) the senotherapeutic NR triflate improved the NAD+ levels of buffy coat-derived platelet concentrates, but cannot prevent storage lesion for 23 days [193]; (iii) 17-α-estradiol late in life extends lifespan in aging UM-HET3 heterogeneous male mice, but NRS does not affect lifespan in either sex [194].
In this regard, NAD+ supplementation (NR, NAM, NA) exhibited emerging roles in replicative and chronological aging in fungi and mammals. For example: (i) Saccharomyces cerevisiae YOR071C gene encodes the high-affinity NR transporter Nrt1 polypeptide [195]; (ii) S. cerevisiae unicellular organism probably represents one of the most recognized experimental aging models for the study of replicative lifespan (RLS, proliferating cells) and chronological lifespan (CLS, non-proliferating cells) [196,197,198].

4.2. Brain Aging, Cognitive Impairment, and Neurodegenerative Diseases

Supplements of NAD+ precursors (NR, NRH, NMN) are a potential way to prevent cognitive decline within aging-associated diseases, such as neurodegenerative disorders [199,200,201]. For example, in models of murine dementia, NRS decreased neuroinflammation, DNA damage, and apoptosis while contributing to maintaining synaptic plasticity, integrity of the blood–brain barrier (BBB), and gut microbiota functionality. It also improved hippocampal synaptic plasticity, learning, and memory in AD. Aβ forming in the brain can be prevented by NRS partly through the upregulation of ubiquitination and proteasomal degradation of PGC-1α-mediated beta-secretase 1 (BACE1) [202,203].
These studies have extended for over a decade. For example, in Tg2576 mice, NR dietary supplementation (250 mg/kg/day) for three months significantly alleviated cognitive damage, increased the NAD+/NADH ratio in the cerebral cortex, and reduced Aβ production [26]. A further study examines DNA repair-deficient 3xTgAD/Polβ+/− mice that exhibited cognitive impairment, synaptic dysfunction, phosphorylated tau (p-tau) pathologies, and neuronal death, the main characteristics of human AD. Here, NRS decreased neuroinflammation, DNA damage, and hippocampal neurons’ apoptosis, increased SIRT3 activity in the brain, improved cognitive function, and restored hippocampal synaptic plasticity [204]. NRS (2.5 g/kg in food for three months) in APP/PS1 transgenic AD and aged mice inhibited serum NAMPT elevation, astrocyte activation, neuroinflammation, senescence, Aβ accumulation, and astrocyte migration to Aβ, as well as improving locomotor activity, cognitive function, behavior, and dementia progression [205,206]. Inhibition of CD38 and NRS alleviated LPS-induced microglial and astrocytic neuroinflammation/neurodegeneration by increasing NAD+ levels in mice brains and by suppressing the NF-κB signaling pathway at the microglia level [207]. In APP/PS1 transgenic (AD) mice, NRS for eight weeks normalized gut dysbiosis for Adlercreutzia, Akkermansia, Bacteroides, Bifidobacterium, Butyricicoccus, Desulfovibrio, Lactobacillus, Olsenella, and Oscillospira microbiota species [208]. In late-onset AD patients, NR and caffeine co-administration partially restores diminished NAD+ availability but does not alter bioenergetic metabolism [209]. In a randomized, double-blinded, PLA-controlled, phase II clinical trial, combined metabolic activators (CMAs) administered in a single dose during the first 28 days and twice daily between days 28 and 84 significantly increased the cognitive capacity and alleviated NAD+ plasma levels and GSH metabolism of AD patients. CMA complex included 12.35 g L-serine (61.75%), 3.73 g L-carnitine tartrate (18.65%), 2.55 g N-acetyl-L-cysteine (12.75%), and 1 g NR (5%) [210].
NR food supplementation for 28 days rescues Ang II-induced cerebral small vessel disease (CSVD) in C57BL/6 mice. NRS significantly reduced glial activation, neuroinflammation, and white matter injury that is associated with cognitive dysfunction. It also supported BBB integrity and vascular remodeling, and improved Ang II-induced CSVD [211].
Further, treatment with NAD+ precursor NRS rescues MT defects in induced pluripotent stem cells (iPSCs) and aging-associated dopaminergic neuronal loss and motor decline and Drosophila models of glucocerebrosidase (GBA)-related Parkinson’s disease (PD) [212]. In a randomized, PLA-controlled, phase I clinical trial in 30 newly diagnosed, treatment-naïve PD patients (NADPARK study), oral NRS 1 g for 30 days significantly increased the level of NAD+ and its related metabolites in the CSF and decreased the inflammatory cytokines amounts also in serum and CSF [213]. In a recent randomized, PLA-controlled crossover trial in 22 healthy older adults using oral NRS (500 mg, twice daily, six weeks), the levels of NAD+ in neuronal-origin enriched plasma extracellular vesicles (NEVs) were increased, and levels of kinases (Aβ42, pJNK, pERK1/2) implicated in neuroinflammation and insulin resistance pathways were inhibited [214]. Moreover, in a double-blind, PLA-controlled trial of 29 PD patients, including cases with common pathogenic mutations in the methylenetetrahydrofolate reductase (MTHFR) gene, it was found that high-dose NRS for 30 days was not associated with altered DNA methylation homeostasis [215].

4.3. Aging and Cancer

NAD+ metabolism is a major feature of cancer pathogenesis (tumorigenesis), being closely related to genome integrity provided by efficient redox homeostasis, MT metabolism, and signal transduction [23]. The anti-tumoral effect of NR was studied in experimental models of hepatocellular carcinoma (HCC), such as the subcutaneous transplantation of tumors in Balb/c nude mice (xenograft) and C57BL/6J mice (allograft) and hematogenous metastatic tumor in nude mice. Daily administration of NR (400 mg/kg) by oral gavage extended the overall survival of HCC mice and decreased the size of allografted tumors and lung, liver, and bone metastases. Also, in vitro TGF-β-induced migration/invasion of HepG2 cells was inhibited by NRS [216].
In a mouse model of C26 adenocarcinoma, NR pellet dietary supplementation significantly improved cancer cachexia and inflammation through the inhibition of specific molecular markers, such as IL-6, TNF-α, PCG-1α, and muscle-specific ubiquitin-proteasome ligases (e.g., mitofusin-2, atrogin-1, muscle RING-finger protein-1 (MuRF-1)) [217]. Oral administration of 200 mg/kg NR in female tumor-bearing rats in a preclinical model of mammary gland cancer induced by N-methyl-nitrosourea (MNU) and previously treated with PTX i.v. injections (three doses of 6.6 mg/kg) led to a decrease in tumor growth and the Ki67 index of tumoral cells and to an improvement in peripheral neuropathy symptoms [218]. In high-risk skin cancer patients, boosting NAD+ with NR, MNM, and NAM p.o. supplements decreased the incidence of keratinocyte carcinoma. This presumably occurred through its cellular protective effects, mainly targeting DNA repair and prevention of possible activation of oncogenic mutations [219].

5. Safety and Bioavailability of Nicotinamide Riboside

5.1. Nicotinamide Riboside Safety

The safety of NRCl was assessed by biological assays, including bacterial reverse mutagenesis tests (Ames test), in vitro chromosome aberration tests, in vivo micronucleus tests, and toxicological studies in rats. These covered acute (24 h), sub-acute (14-day), and chronic (90-day) administration. Genotoxicity of NRCl was not observed, and lethality was absent even at a dose of 5000 mg/kg (p.o.). NRCl had the lowest-observed-adverse-effect level (LOAEL) and the no-observed-adverse-effect level (NOAEL) at the dose of 1000 mg/kg/day and 300 mg/kg/day, respectively. Here, the liver, kidney, ovary, and testicles were the target organs for oral toxicity assays [220].
A randomized, double-blind, PLA-controlled clinical trial for eight weeks was performed in overweight healthy women and men who received different doses of NRCl (orally 100 mg, 300 mg, 1000 mg). Between NRCl and PLA groups or between different NRCl groups, neither important differences nor adverse effects were observed. However, within two weeks of NRCl supplementation, significant increases in serum levels of NAD+ and NAD+ metabolites were recorded [3].
In a randomized, double-blind, PLA-controlled study, the safety and efficacy of NRPT (Basis™) were assessed for eight weeks in 120 healthy adult volunteers (60–80 years old). In addition to the significant increase in NAD+ blood levels, repeated doses of NRPT supplementation did not cause any adverse effects [221]. The European Food Safety Authority (EFSA) Panel on Nutrition, Novel Foods and Allergens (NDA) of the European Commission (EC) concluded that NR and NRCl are “safe for the healthy adult population, excluding pregnant and lactating women, and that an intake of up to 230 mg/day is safe for pregnant and lactating women”. Moreover, for the two active compounds, there are no concerns about genotoxicity and their safety/tolerability in human studies [222,223,224]. Three more leading international regulatory authorities, including the U.S. Food and Drug Administration (FDA), Health Canada (HC), and the Therapeutic Goods Administration (TGA) of Australia, concluded that NR is safer for human use than other NAD+ metabolites [225].
Similar studies were performed on NR-related compounds. A non-randomized study of single oral supplementation within the daily tolerable upper levels of NAM affects blood NAM and NAD+ amounts in six healthy humans. NAM daily intake of up to 200 mg was found, at 0.5 h, a 30-fold increase in its maximum amount in the whole blood of male volunteers. At 12 h, NAD+ serum concentrations have reached the maximum level. Metabolomic variations due to NAM dietary supplementation lasted one day, and then after two days, they returned to the baseline [226]. Moreover, in another study, oral supplementation of NAM (500 mg) significantly increased blood NAD+ levels after 12 h in a cohort of five healthy adult subjects [227].
A toxicological study assessed the safety of Restorin® NMN, a direct NAD+ precursor (β-NMN) in a high-purity formulation, by oral administration at 500, 1000, and 2000 mg/kg/day in Sprague–Dawley rats over three months followed by a recovery period of two weeks. No adverse effects were observed at the dose of 500 mg/kg/day. Restorin® NMN has a LOAEL of 2000 mg/kg/day and a NOAEL of 500 mg/kg/day in male rats and 1000 mg/kg/day in female rats [228].
In a double-blind, PLA-controlled, randomized clinical trial, oral administration of NMN (250 mg/day for 12 weeks) proved to be safe by increasing the level of blood NAD+ in 30 healthy volunteers. A significant increase in the level of NA mononucleotide (NAMN), rather than in the level of NMN, was also seen [229]. MIB-626, an oral formulation of a microcrystalline unique polymorph of β-NMN, increased circulating NAD+ levels in a double-blind, PLA-controlled trial in 32 overweight or obese middle-aged and older adults (55–80 years). Once or twice daily, 1000 mg MIB-626 supplementation proved to be safe and well tolerated. Age, sex, or BMI did not influence the variations in serum amounts of NMN or NAD+ [230].

5.2. Nicotinamide Riboside Bioavailability

In mice and humans, NR is the most orally bioavailable of NAD+ precursors. Amounts of NAD+ in mouse liver were elevated by oral NRS with better pharmacokinetics (PK) than seen with NA and NAM. In one healthy volunteer, a pilot study showed that a single dose of oral NRS elevated serum NAD+ levels as much as 2.7-fold. The first clinical trial of NR PK in healthy humans highlighted that 100, 300, and 1000 mg of NR single dose increased the NAD+ metabolome in a dose-dependent manner [231].
An open-label, non-randomized study of the PK of the NR nutritional supplement and its effects on blood NAD+ levels was performed in eight healthy volunteers. Oral NR (250 mg) was administered daily for the first two days. Then, the maximum dose was increased to 1000 mg twice daily on days 7 and 8. NRS did not induce adverse effects and was well tolerated. In NR and NAD+ levels, starting from baseline to the ninth day, the absolute changes were highly correlated [232,233].
BM stromal cell antigen 1 (BST1) regulates NR metabolism via its dual roles in glycohydrolase and base-exchange enzyme activities. Two different pathways contribute to the increase in NAD+ level after oral NR treatment: (i) NR salvage pathway in the early phase of NR direct absorption, and (ii) Preiss–Handler pathway in the late phase of NR hydrolysis to NAM by BST1 followed by transformation to NA by the distal gut microbiome [234,235].
Pre-steady-state and steady-state kinetic analyses were made for the N-ribosyl hydrolase activity of hCD157. Cell-surface hCD157 binds and catalyzes the slow hydrolysis of NR and NA riboside (NAR) after the formation of an initial complex with NAD+ derivatives [236].
The import of NR and NAR into human cells is mediated by the equilibrative nucleoside transporter (ENT) family of proteins, ENT1, ENT2, and ENT4. After entering the HEK293 cells, NR is activated by NRK phosphorylation in the cytosol and then transformed to yield NAM [237]. NMN and NR metabolism in mammalian cells is predominantly controlled by NRK1 and purine nucleoside phosphorylase (PNP). Simultaneous PNP downregulation could improve the health advantages of NRS [14,238].
Kulikova et al. (2015) studied the generation, release, and uptake of the NAD+ precursor NAR by human cells. Under normal culture conditions, untransfected HeLa cells generated and released a sufficient level of NAR and NR as whole compounds derived from NAD+ metabolism [239].
A pilot study investigated changes in the human plasma and urine NAD+ metabolome during a 6 h i.v. infusion of 3 μM/min NAD+. No changes were seen after two hours in the levels of plasma NAD+ or its metabolites, such as NAM, MeNAM, adenosine diphosphate ribose (ADPR), and NMN. After six hours, increased urinary excretion of NAD+ and MeNAM was seen; at the same time, NAM urinary level was not increased [240].
In a simulated intestinal fluid using porcine pancreatin, NR trioleate chloride, a newly synthesized hydrophobic NR derivative, was partly digestible and easily released NR [241].
NRH is a potent enhancer of NAD+ levels in vitro in mammalian cells, significantly increasing NAD+/NADH ratio, as well as in vivo after i.p. injection to C57BL/6J mice [242]. Although there are minor structural dissimilarities, NRH defines a new pathway for NAD+ biosynthesis, independent of NRK1, and acts as an orally bioavailable NAD+ precursor also in mammalian cells and mouse tissues [243]. In a recent study, it was found that NR and dihydronicotinic acid riboside synergistically act to increase NAD+ intracellular levels by generating NRH [244].

6. Borate-Stabilized Nicotinamide Riboside

This collection of research results is cited to encourage the consumption of NR as a dietary supplement. Indeed, several companies now offer NR in pill form for this purpose. Unfortunately, NR is a reactive species, often unstable during manufacture, transport, and storage. Accordingly, HPLC analysis of many of the commercially available products found them to not be pure NR but rather mixtures of NR and various other materials. Often, the NR is present in small amounts. These facts led us to develop NR borate and less-expensive routes to make it. These are reviewed below.

6.1. Prebiotic Synthesis

While this is irrelevant to the use of NR as a dietary supplement, it is of conceptual interest to note that the synthesis of NR borate was grounded in considerations of prebiotic chemistry. The analysis began with the observation that the chemical and biological syntheses of NR are extremely challenging, with the challenge centered on the bond that joins the NAM ring to the ribose ring. Thus, while several methods of chemical and biological syntheses of NR are reported [245], these are rather expensive. At the time of this writing, only one U.S. company (ChromaDex) produced NRCl [3,220]. Further, NR is unstable. When analyzed by HPLC and other standard analytical methods, many samples of commercial NR are, in fact, mixtures of products that indicate the decomposition of the material [246].
Prebiotic chemistry was applied to this problem. Here, Kim and Benner discovered that NR phosphate emerges in stable form by direct reaction of ribose-1,2-cyclic phosphate with NAM [247,248]. Ribose-1,2-cyclic phosphate is available inexpensively from ribose and amidotriphosphate, which is available from the very inexpensive cyclic trimetaphosphate and ammonia. The phosphorylated NR product is then enzymatically dephosphorylated by a phosphatase [249]. In the final stage, NR is stabilized by borate, resulting in NR borate [250]. Borate-stabilized NR was found to have high stability against thermal and chemical degradation [250,251]. This mitigates problems in current commercial NR.
The borate might be seen as simply a stabilizing species. However, recent research suggests that borate itself might be a key element in NAD+ metabolism [252,253]. Borate released by the ribose–borate complex by hydrolysis could be a reservoir of physiological boron (B). In addition, borate esters can remain undigested in the upper gastric system and thus be more accessible to the microbiota [254]. According to new insights into the essentiality of B in the healthy symbiosis between the microbiota and the human host, NR borate becomes an interesting prebiotic that delivers B that is essential to bacteria for the synthesis of autoinducer-2 (AI-2) and for strengthening the colonic mucus [254,255,256].
The prebiotic synthesis and borate stabilization of NR had two very important advantages: (i) it is more cost-effective due to less expensive starting materials; (ii) NR is more stable, mitigating the fact that many samples of commercial NRCl are mixtures of products that indicate decomposition of the material (Figure 1).

6.2. Solubility and Degradation Kinetics of Nicotinamide Riboside Borate

The possibility that NR borate provides, at last, a stable form of NR should motivate the study of its effect in preclinical and clinical studies. Since borate esters are dynamic, with borate dissociating from NR upon dilution, we expect that the effects of NR borate should closely parallel the effects of NR.
The solubility and degradation kinetics of NR borate were examined in various physiological media, including 0.1 N hydrochloric acid (pH 1.5), sodium acetate buffer (pH 5.0), water (pH 7.0), and phosphate-buffered saline (pH 7.4), using HPLC analysis. Furthermore, stability studies were conducted to investigate how different pH levels and temperatures influenced the degradation kinetics of NR borate. The results revealed significant variations in the solubility and stability of NR borate within the tested biological solutions. The solubility of NR borate was found to be pH-dependent, increasing as the pH level rose [257].
Using the shake-flask method at room temperature, the solubility of NR borate was evaluated. The measurements showed that NR borate had solubilities of 1972.7 ± 15.4 mg/mL, 1060.5 ± 31.0 mg/mL, and 926.0 ± 34.4 mg/mL at pH 1.5, 5.0, and 7.4, respectively. These findings indicate that NR borate demonstrated high solubility in all three solutions with varying pH levels, meeting the standards outlined in the U.S. Pharmacopoeia recommendations [257] (Figure 2).
To investigate the spontaneity of NR borate degradation in various pH solutions, the Gibbs free energy (ΔG) was calculated. The ΔG value obtained for NR borate was 2.43 kcal/mol [257].
When examining NR borate as a nutritional supplement, it is important to investigate the generation of NAM and the ribose–borate complex, which acts as a physiological reservoir of B, as degradation byproducts. It is understood that the existence of anionic borate can impede the conversion of NR into NAM and trigger the activation of SIRT. While there is limited scientific data on the slight degradation of NR in simulated gastric fluids, substantial degradation seems to occur in the intestine due to its elevated pH levels [233]. The stability of NR borate is significantly influenced by neutral conditions, exemplified by the pH 7.4 employed in this study. In the Arrhenius model of NR borate degradation, the rate of degradation is higher at pH 7.4 compared to pH 1.5 and pH 5.0, supporting the notion that the degradation of NR borate is influenced by the concentration of HO ions in the solution. It is crucial to exercise caution when using an aqueous solution of NR borate, particularly at high pH values where an excess of HO ions can accelerate the degradation process due to hydrolysis. Our findings suggest that there may be a slightly distinct mechanism of degradation for NR and NR borate under acidic conditions compared to neutral or alkaline conditions. Consequently, this model can serve as a general reference for comprehending the stability of NR borate in various buffered solutions. Additionally, temperature plays a role in the content of NR borate in solution, with a 10 °C increase approximately doubling its degradation rate under any pH condition.
The estimation of logP is crucial for predicting the permeability of NR borate. It is well established that compounds with estimated logP values less than 1.72 exhibit low permeability. According to the Biopharmaceutical Classification System (BCS), NR borate is classified as Class III due to its high solubility but anticipated low permeability, as indicated by a logP value of −4.17. This physiological characteristic may be linked to the proportion of released degradation products. Ultra-high-performance liquid chromatography (UHPLC) spectra revealed that the degradation products of NR borate in aqueous solutions are NAM and the ribose–borate complex. Consequently, the concentration of NAM can serve as a marker for the degradation of both NR borate and NRCl, the commercially available form. In this study, a developed high-performance thin-layer chromatography (HPTLC) method was employed to separate and quantify NR and NAM peaks with a remarkable separation resolution (R = 1.8), surpassing the acceptable resolution (R > 1) recommended by the U.S. FDA [257]. This method offers a rapid means of detecting NR, its degradation product NAM, and the ribose–borate complex. Our research data present, for the first time, a definitive upper-temperature limit for the processing of NR borate and its products, which can impact the production of supplements at various stages. The optimization of NR borate administration will have a significant influence on its efficacy. Our study aimed to uncover the mechanism by which NR borate degradation produces a physiological product when exposed to simulated physiological fluids in vitro. The development of an industrial pilot NR borate opens up possibilities for further research, which could have a profound impact on its potential therapeutic applications. Consequently, under these conditions, NR borate solutions and formulations have the potential for high stability. The findings of this study will prove valuable for pharmaceutical scientists involved in manipulating NR borate working solutions and designing appropriate formulations for NR borate delivery. We observed a clear dependence of NR borate solubility on pH. NR borate has exhibited extreme instability at pH levels greater than 8.0 while remaining stable in both acidic and neutral environments.
As a result, NR borate may offer advantages over NRCl as a nutritional supplement, primarily due to the boric acid (BA) residue’s ability to block the glycosidic bond between the pyrimidine base and ribose. The study has shed light on the key physicochemical properties that should be considered when utilizing NR borate in both in vitro and in vivo experiments. The described HPTLC method has the potential for distinguishing between NR borate and its degradation products in environments with pH conditions similar to those found in the GI tract. The obtained data also hold significance for the processing, production, and storage of this supplement.
In addition, the greatest direction of the use of NR borate could be as a prebiotic compound. The latest research shows that the presence of NR in the colon has a very high nutritional value, and through its ester form of BA, it can become a nutraceutical with a role in healthy symbiosis as an essential B element as well as an essential nutrient in the health metabolism of the human microbiome [143,144]. In vivo, B organic compounds have been shown to be less toxic than BA and inorganic borates [253].

7. Conclusions and Perspectives

NAD+ is known classically as a metabolite that stands astride both catabolic and anabolic pathways throughout the metabolism that is taught in introductory biochemistry courses. However, non-classical studies starting over a decade ago found that it is also involved in higher-order functions, in part because of its involvement in the activation of SIRTs and the support of the MT unfolded protein response. Many studies have suggested that NAD+ is involved in an extensive spectrum of pathologies, including neurodegenerative disorders, cardiomyopathy, obesity, and diabetes. Further, healthy aging and longevity appear to be closely related to NAD+ and its related metabolites, including through NR and NMN. This system appears to have prophylactic and therapeutic value in improving age-associated neurodegenerative, CV, and metabolic diseases and conditions.
Accordingly, many are now recommending the consumption of materials in this system as dietary supplements, hoping to achieve overall improvements in human health. Among NAD+ precursors, NR appears to have special values. These include better tolerance, better uptake, and overall greater potency.
Unfortunately, NR is a reactive molecule, often unstable during its manufacturing, transport, and storage. Indeed, HPLC analyses of many commercial samples of NR show that they contain substantial amounts of material that are not, in fact, NR. More stable derivatives of NR that are easily converted upon consumption into NR are therefore desired.
Recently work related to prebiotic chemistry provided the borate derivative of NR. NR borate is considerably more stable than NR itself. However, immediately upon consumption, the borate dissociates from the NR borate and is lost in the body through dilution and binding to other species, notably carbohydrates such as fructose and glucose. The NR left behind is expected to behave pharmacologically in ways identical to NR itself.
This review provides a comprehensive uncritical summary through Q1 of 2023 of the literature that makes the case for the consumption of NR. It then summarizes the challenges of delivering quality NR to consumers using standard synthesis, manufacture, shipping, and storage approaches. It concludes by outlining the advantages of NR borate in these processes.

Author Contributions

Conceptualization, A.B., I.R.S., G.D.M. and S.A.B.; methodology, M.V.C., O.E.N., A.S.P., G.R., A.R. and C.B.; writing—original draft preparation, I.R.S., G.D.M. and L.E.B.; writing—review and editing, A.B., I.R.S., G.D.M. and S.A.B.; supervision, A.B., I.R.S., G.D.M. and S.A.B.; funding acquisition, A.B., I.R.S. and J.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the University of Medicine and Pharmacy of Craiova, project number 26/29c. Preparation of this manuscript was also supported in part by the US National Science Foundation EAR-2213438.

Institutional Review Board Statement

The manuscript does not contain experiments on laboratory animals.

Informed Consent Statement

The manuscript does not contain clinical studies or patient data.

Data Availability Statement

Data described in the manuscript will be made publicly and freely available without restriction at: https://drive.google.com/drive/folders/1BraUbHcRZ4vNJfuEIAzlzwsd-o4BSxMX?usp=sharing (accessed on 20 June 2023).

Conflicts of Interest

The authors declare no conflict of interest. The funder had no role in the design of the study; in the collection, analysis, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Penberthy, W.T.; Kirkland, J.B. Chapter 19: Niacin. In Present Knowledge in Nutrition, 10th ed.; Erdman, J.W., Jr., MacDonald, I.A., Zeisel, S.H., Eds.; International Life Sciences Institute (ILSI): Ames, IA, USA; Wiley–Blackwell, John Wiley & Sons: Ames, IA, USA, 2012; pp. 293–306. [Google Scholar] [CrossRef]
  2. Dollerup, O.L.; Christensen, B.; Svart, M.; Schmidt, M.S.; Sulek, K.; Ringgaard, S.; Stødkilde-Jørgensen, H.; Møller, N.; Brenner, C.; Treebak, J.T.; et al. A randomized placebo-controlled clinical trial of nicotinamide riboside in obese men: Safety, insulin-sensitivity, and lipid-mobilizing effects. Am. J. Clin. Nutr. 2018, 108, 343–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Conze, D.; Brenner, C.; Kruger, C.L. Safety and metabolism of long-term administration of NIAGEN (nicotinamide riboside chloride) in a randomized, double-blind, placebo-controlled clinical trial of healthy overweight adults. Sci. Rep. 2019, 9, 9772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Mehmel, M.; Jovanović, N.; Spitz, U. Nicotinamide riboside—The current state of research and therapeutic uses. Nutrients 2020, 12, 1616. [Google Scholar] [CrossRef]
  5. Gilbert, W. Origin of life: The RNA world. Nature 1986, 319, 618. [Google Scholar] [CrossRef]
  6. White, H.B., 3rd. Coenzymes as fossils of an earlier metabolic state. J. Mol. Evol. 1976, 7, 101–104. [Google Scholar] [CrossRef] [PubMed]
  7. Chi, Y.; Sauve, A.A. Nicotinamide riboside, a trace nutrient in foods, is a vitamin B3 with effects on energy metabolism and neuroprotection. Curr. Opin. Clin. Nutr. Metab. Care 2013, 16, 657–661. [Google Scholar] [CrossRef]
  8. Martens, C.R.; Denman, B.A.; Mazzo, M.R.; Armstrong, M.L.; Reisdorph, N.; McQueen, M.B.; Chonchol, M.; Seals, D.R. Chronic nicotinamide riboside supplementation is well-tolerated and elevates NAD+ in healthy middle-aged and older adults. Nat. Commun. 2018, 9, 1286. [Google Scholar] [CrossRef]
  9. Trammell, S.A.J.; Yu, L.; Redpath, P.; Migaud, M.E.; Brenner, C. Nicotinamide riboside is a major NAD+ precursor vitamin in cow milk. J. Nutr. 2016, 146, 957–963. [Google Scholar] [CrossRef] [Green Version]
  10. Lee, H.J.; Yang, S.J. Supplementation with nicotinamide riboside reduces brain inflammation and improves cognitive function in diabetic mice. Int. J. Mol. Sci. 2019, 20, 4196. [Google Scholar] [CrossRef] [Green Version]
  11. Bogan, K.L.; Brenner, C. Nicotinic acid, nicotinamide, and nicotinamide riboside: A molecular evaluation of NAD+ precursor vitamins in human nutrition. Annu. Rev. Nutr. 2008, 28, 115–130. [Google Scholar] [CrossRef] [Green Version]
  12. Belenky, P.; Stebbins, R.; Bogan, K.L.; Evans, C.R.; Brenner, C. Nrt1 and Tna1-independent export of NAD+ precursor vitamins promotes NAD+ homeostasis and allows engineering of vitamin production. PLoS ONE 2011, 6, e19710. [Google Scholar] [CrossRef] [Green Version]
  13. Bieganowski, P.; Brenner, C. Discoveries of nicotinamide riboside as a nutrient and conserved NRK genes establish a Preiss–Handler independent route to NAD+ in fungi and humans. Cell 2004, 117, 495–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Ratajczak, J.; Joffraud, M.; Trammell, S.A.; Ras, R.; Canela, N.; Boutant, M.; Kulkarni, S.S.; Rodrigues, M.; Redpath, P.; Migaud, M.E.; et al. NRK1 controls nicotinamide mononucleotide and nicotinamide riboside metabolism in mammalian cells. Nat. Commun. 2016, 7, 13103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Schmidt, M.S.; Brenner, C. Absence of evidence that Slc12a8 encodes a nicotinamide mononucleotide transporter. Nat. Metab. 2019, 1, 660–661. [Google Scholar] [CrossRef] [PubMed]
  16. Canto, C. NAD+ precursors: A questionable redundancy. Metabolites 2022, 12, 630. [Google Scholar] [CrossRef] [PubMed]
  17. Cantó, C.; Auwerx, J. NAD+ as a signaling molecule modulating metabolism. Cold Spring Harb. Symp. Quant. Biol. 2011, 76, 291–298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Imai, S.; Guarente, L. NAD+ and sirtuins in aging and disease. Trends Cell Biol. 2014, 24, 464–471. [Google Scholar] [CrossRef]
  19. Zhang, N.; Sauve, A.A. Regulatory effects of NAD+ metabolic pathways on sirtuin activity. Prog. Mol. Biol. Transl. Sci. 2018, 154, 71–104. [Google Scholar] [CrossRef]
  20. Okabe, K.; Yaku, K.; Tobe, K.; Nakagawa, T. Implications of altered NAD metabolism in metabolic disorders. J. Biomed. Sci. 2019, 26, 34. [Google Scholar] [CrossRef] [Green Version]
  21. Buonvicino, D.; Ranieri, G.; Pittelli, M.; Lapucci, A.; Bragliola, S.; Chiarugi, A. SIRT1-dependent restoration of NAD+ homeostasis after increased extracellular NAD+ exposure. J. Biol. Chem. 2021, 297, 100855. [Google Scholar] [CrossRef]
  22. Carpi, F.M.; Cortese, M.; Orsomando, G.; Polzonetti, V.; Vincenzetti, S.; Moreschini, B.; Coleman, M.; Magni, G.; Pucciarelli, S. Simultaneous quantification of nicotinamide mononucleotide and related pyridine compounds in mouse tissues by UHPLC–MS/MS. Sep. Sci. Plus 2018, 1, 22–30. [Google Scholar] [CrossRef]
  23. Demarest, T.G.; Babbar, M.; Okur, M.N.; Dan, X.; Croteau, D.L.; Fakouri, N.B.; Mattson, M.P.; Bohr, V.A. NAD+ metabolism in aging and cancer. Annu. Rev. Cancer Biol. 2019, 3, 105–130. [Google Scholar] [CrossRef]
  24. Yang, T.; Chan, N.Y.K.; Sauve, A.A. Syntheses of nicotinamide riboside and derivatives: Effective agents for increasing nicotinamide adenine dinucleotide concentrations in mammalian cells. J. Med. Chem. 2007, 50, 6458–6461. [Google Scholar] [CrossRef]
  25. Cantó, C.; Houtkooper, R.H.; Pirinen, E.; Youn, D.Y.; Oosterveer, M.H.; Cen, Y.; Fernandez-Marcos, P.J.; Yamamoto, H.; Andreux, P.A.; Cettour-Rose, P.; et al. The NAD+ precursor nicotinamide riboside enhances oxidative metabolism and protects against high-fat diet-induced obesity. Cell Metab. 2012, 15, 838–847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Gong, B.; Pan, Y.; Vempati, P.; Zhao, W.; Knable, L.; Ho, L.; Wang, J.; Sastre, M.; Ono, K.; Sauve, A.A.; et al. Nicotinamide riboside restores cognition through an upregulation of proliferator-activated receptor-γ coactivator 1α regulated β-secretase 1 degradation and mitochondrial gene expression in Alzheimer’s mouse models. Neurobiol. Aging 2013, 34, 1581–1588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Cantó, C.; Menzies, K.J.; Auwerx, J. NAD+ metabolism and the control of energy homeostasis: A balancing act between mitochondria and the nucleus. Cell Metab. 2015, 22, 31–53. [Google Scholar] [CrossRef] [Green Version]
  28. Conlon, N.; Ford, D. A systems-approach to NAD+ restoration. Biochem. Pharmacol. 2022, 198, 114946. [Google Scholar] [CrossRef]
  29. Sharma, C.; Donu, D.; Cen, Y. Emerging role of nicotinamide riboside in health and diseases. Nutrients 2022, 14, 3889. [Google Scholar] [CrossRef]
  30. Helman, T.; Braidy, N. Importance of NAD+ anabolism in metabolic, cardiovascular and neurodegenerative disorders. Drugs Aging 2023, 40, 33–48. [Google Scholar] [CrossRef]
  31. Fang, J.; Wu, H.; Zhang, J.; Mao, S.; Shi, H.; Yu, D.; Chen, Z.; Su, K.; Xing, Y.; Dong, H.; et al. A reduced form of nicotinamide riboside protects the cochlea against aminoglycoside-induced ototoxicity by SIRT1 activation. Biomed. Pharmacother. 2022, 150, 113071. [Google Scholar] [CrossRef]
  32. She, J.; Sheng, R.; Qin, Z.H. Pharmacology and potential implications of nicotinamide adenine dinucleotide precursors. Aging Dis. 2021, 12, 1879–1897. [Google Scholar] [CrossRef] [PubMed]
  33. Belenky, P.; Bogan, K.L.; Brenner, C. NAD+ metabolism in health and disease. Trends Biochem. Sci. 2007, 32, 12–19. [Google Scholar] [CrossRef] [PubMed]
  34. Belenky, P.; Racette, F.G.; Bogan, K.L.; McClure, J.M.; Smith, J.S.; Brenner, C. Nicotinamide riboside promotes Sir2 silencing and extends lifespan via Nrk and Urh1/Pnp1/Meu1 pathways to NAD+. Cell 2007, 129, 473–484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Yoshino, J.; Baur, J.A.; Imai, S.I. NAD+ intermediates: The biology and therapeutic potential of NMN and NR. Cell Metab. 2018, 27, 513–528. [Google Scholar] [CrossRef] [Green Version]
  36. Stock, A.J.; Ayyar, S.; Kashyap, A.; Wang, Y.; Yanai, H.; Starost, M.F.; Tanaka-Yano, M.; Bodogai, M.; Sun, C.; Wang, Y.; et al. Boosting NAD ameliorates hematopoietic impairment linked to short telomeres in vivo. Geroscience 2023, 2023, 1–16. [Google Scholar] [CrossRef]
  37. Hong, G.; Zheng, D.; Zhang, L.; Ni, R.; Wang, G.; Fan, G.C.; Lu, Z.; Peng, T. Administration of nicotinamide riboside prevents oxidative stress and organ injury in sepsis. Free Radic. Biol. Med. 2018, 123, 125–137. [Google Scholar] [CrossRef]
  38. Wu, J.; Singh, K.; Lin, A.; Meadows, A.M.; Wu, K.; Shing, V.; Bley, M.; Hassanzadeh, S.; Huffstutler, R.D.; Schmidt, M.S.; et al. Boosting NAD+ blunts TLR4-induced type I IFN in control and systemic lupus erythematosus monocytes. J. Clin. Invest. 2022, 132, e139828. [Google Scholar] [CrossRef]
  39. Zhang, X.; Tian, B.; Deng, Q.; Cao, J.; Ding, X.; Liu, Q.; Zhang, Y.; Ye, C.; Deng, C.; Qiu, L.; et al. Nicotinamide riboside relieves the severity of experimental necrotizing enterocolitis by regulating endothelial function via eNOS deacetylation. Free Radic. Biol. Med. 2022, 184, 218–229. [Google Scholar] [CrossRef]
  40. Fan, X.K.; Xu, C.Q.; Cao, K.Q.; Zhao, G.J.; Hong, G.L.; Lu, Z.Q. Study on the protective effect and mechanism of nicotinamide riboside on lung injury in paraquat intoxicated mice. Zhonghua Lao Dong Wei Sheng Zhi Ye Bing Za Zhi 2022, 40, 561–567. [Google Scholar] [CrossRef]
  41. Selli, J.; Vural Keles, D.; Keles, O.N.; Celik, M.; Yetim, Z. Nicotinamide riboside preserves ovarian injury in experimental sepsis model in rats. Eurasian J. Med. 2023, 55, 128–134. [Google Scholar] [CrossRef]
  42. DiNicolantonio, J.J.; McCarty, M.F.; O’Keefe, J.H. Nutraceutical activation of Sirt1: A review. Open Heart 2022, 9, e002171. [Google Scholar] [CrossRef] [PubMed]
  43. Cheng, Y.H.; Zhao, J.H.; Zong, W.F.; Wei, X.J.; Xu, Z.; Yuan, Y.; Jiang, Y.F.; Luo, X.; Wang, W.; Qu, W.S. Acute treatment with nicotinamide riboside chloride reduces hippocampal damage and preserves the cognitive function of mice with ischemic injury. Neurochem. Res. 2022, 47, 2244–2253. [Google Scholar] [CrossRef] [PubMed]
  44. Lenherr, N.; Christodoulou, J.; Duley, J.; Dobritzsch, D.; Fairbanks, L.; Datta, A.N.; Filges, I.; Gürtler, N.; Roelofsen, J.; van Kuilenburg, A.B.P.; et al. Co-therapy with S-adenosylmethionine and nicotinamide riboside improves T-cell survival and function in Arts syndrome (PRPS1 deficiency). Mol. Genet. Metab. Rep. 2021, 26, 100709. [Google Scholar] [CrossRef] [PubMed]
  45. Kodali, M.; Jankay, T.; Shetty, A.K.; Reddy, D.S. Pathophysiological basis and promise of experimental therapies for Gulf War Illness, a chronic neuropsychiatric syndrome in veterans. Psychopharmacology 2023, 240, 673–697. [Google Scholar] [CrossRef]
  46. Joshi, U.; Evans, J.E.; Pearson, A.; Saltiel, N.; Cseresznye, A.; Darcey, T.; Ojo, J.; Keegan, A.P.; Oberlin, S.; Mouzon, B.; et al. Targeting sirtuin activity with nicotinamide riboside reduces neuroinflammation in a GWI mouse model. Neurotoxicology 2020, 79, 84–94. [Google Scholar] [CrossRef]
  47. Jiang, Y.; Liu, Y.; Gao, M.; Xue, M.; Wang, Z.; Liang, H. Nicotinamide riboside alleviates alcohol-induced depression-like behaviours in C57BL/6J mice by altering the intestinal microbiota associated with microglial activation and BDNF expression. Food Funct. 2020, 11, 378–391. [Google Scholar] [CrossRef] [PubMed]
  48. Vaur, P.; Brugg, B.; Mericskay, M.; Li, Z.; Schmidt, M.S.; Vivien, D.; Orset, C.; Jacotot, E.; Brenner, C.; Duplus, E. Nicotinamide riboside, a form of vitamin B3, protects against excitotoxicity-induced axonal degeneration. FASEB J. 2017, 31, 5440–5452. [Google Scholar] [CrossRef] [Green Version]
  49. Kitaoka, Y.; Sase, K.; Tsukahara, C.; Fujita, N.; Arizono, I.; Takagi, H. Axonal protection by nicotinamide riboside via SIRT1-autophagy pathway in TNF-induced optic nerve degeneration. Mol. Neurobiol. 2020, 57, 4952–4960. [Google Scholar] [CrossRef]
  50. Chen, A.; Kristiansen, C.K.; Hong, Y.; Kianian, A.; Fang, E.F.; Sullivan, G.J.; Wang, J.; Li, X.; Bindoff, L.A.; Liang, K.X. Nicotinamide riboside and metformin ameliorate mitophagy defect in induced pluripotent stem cell-derived astrocytes with POLG mutations. Front. Cell Dev. Biol. 2021, 9, 737304. [Google Scholar] [CrossRef]
  51. Sun, C.; Seranova, E.; Cohen, M.A.; Chipara, M.; Roberts, J.; Astuti, D.; Palhegyi, A.M.; Acharjee, A.; Sedlackova, L.; Kataura, T.; et al. NAD depletion mediates cytotoxicity in human neurons with autophagy deficiency. Cell Rep. 2023, 42, 112372. [Google Scholar] [CrossRef]
  52. Yang, B.; Dan, X.; Hou, Y.; Lee, J.H.; Wechter, N.; Krishnamurthy, S.; Kimura, R.; Babbar, M.; Demarest, T.; McDevitt, R.; et al. NAD+ supplementation prevents STING-induced senescence in ataxia telangiectasia by improving mitophagy. Aging Cell 2021, 20, e13329. [Google Scholar] [CrossRef]
  53. Veenhuis, S.J.G.; van Os, N.J.H.; Janssen, A.J.W.M.; van Gerven, M.H.J.C.; Coene, K.L.M.; Engelke, U.F.H.; Wevers, R.A.; Tinnevelt, G.H.; Ter Heine, R.; van de Warrenburg, B.P.C.; et al. Nicotinamide riboside improves ataxia scores and immunoglobulin levels in ataxia telangiectasia. Mov. Disord. 2021, 36, 2951–2957. [Google Scholar] [CrossRef]
  54. Steinbruecker, K.; Tiefenthaler, E.; Schernthaler, E.M.; Jungwirth, J.; Wortmann, S.B. Nicotinamide riboside for ataxia telangiectasia: Report of an early treated individual. Neuropediatrics 2022, 54, 78–81. [Google Scholar] [CrossRef]
  55. Gerasimenko, M.; Cherepanov, S.M.; Furuhara, K.; Lopatina, O.; Salmina, A.B.; Shabalova, A.A.; Tsuji, C.; Yokoyama, S.; Ishihara, K.; Brenner, C.; et al. Nicotinamide riboside supplementation corrects deficits in oxytocin, sociability and anxiety of CD157 mutants in a mouse model of autism spectrum disorder. Sci. Rep. 2020, 10, 10035. [Google Scholar] [CrossRef]
  56. Gerasimenko, M.; Higashida, H. Remission of social behavior impairment by oral administration of a precursor of NAD in CD157, but not in CD38, knockout mice. Front. Immunol. 2023, 14, 1166609. [Google Scholar] [CrossRef] [PubMed]
  57. Harlan, B.A.; Killoy, K.M.; Pehar, M.; Liu, L.; Auwerx, J.; Vargas, M.R. Evaluation of the NAD+ biosynthetic pathway in ALS patients and effect of modulating NAD+ levels in hSOD1-linked ALS mouse models. Exp. Neurol. 2020, 327, 113219. [Google Scholar] [CrossRef] [PubMed]
  58. Zhou, Q.; Zhu, L.; Qiu, W.; Liu, Y.; Yang, F.; Chen, W.; Xu, R. Nicotinamide riboside enhances mitochondrial proteostasis and adult neurogenesis through activation of mitochondrial unfolded protein response signaling in the brain of ALS SOD1G93A mice. Int. J. Biol. Sci. 2020, 16, 284–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Obrador, E.; Salvador, R.; Marchio, P.; López-Blanch, R.; Jihad-Jebbar, A.; Rivera, P.; Vallés, S.L.; Banacloche, S.; Alcácer, J.; Colomer, N.; et al. Nicotinamide riboside and pterostilbene cooperatively delay motor neuron failure in ALS SOD1G93A mice. Mol. Neurobiol. 2021, 58, 1345–1371. [Google Scholar] [CrossRef]
  60. Freeberg, K.A.; Udovich, C.A.C.; Martens, C.R.; Seals, D.R.; Craighead, D.H. Dietary supplementation with NAD+-boosting compounds in humans: Current knowledge and future directions. J. Gerontol. A Biol. Sci. Med. Sci. 2023, glad106. [Google Scholar] [CrossRef]
  61. Brown, K.D.; Maqsood, S.; Huang, J.Y.; Pan, Y.; Harkcom, W.; Li, W.; Sauve, A.; Verdin, E.; Jaffrey, S.R. Activation of SIRT3 by the NAD+ precursor nicotinamide riboside protects from noise-induced hearing loss. Cell Metab. 2014, 20, 1059–1068. [Google Scholar] [CrossRef] [Green Version]
  62. Han, S.; Du, Z.; Liu, K.; Gong, S. Nicotinamide riboside protects noise-induced hearing loss by recovering the hair cell ribbon synapses. Neurosci. Lett. 2020, 725, 134910. [Google Scholar] [CrossRef] [PubMed]
  63. Okur, M.N.; Mao, B.; Kimura, R.; Haraczy, S.; Fitzgerald, T.; Edwards-Hollingsworth, K.; Tian, J.; Osmani, W.; Croteau, D.L.; Kelley, M.W.; et al. Short-term NAD+ supplementation prevents hearing loss in mouse models of Cockayne syndrome. NPJ Aging Mech. Dis. 2020, 6, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Hamity, M.V.; White, S.R.; Walder, R.Y.; Schmidt, M.S.; Brenner, C.; Hammond, D.L. Nicotinamide riboside, a form of vitamin B3 and NAD+ precursor, relieves the nociceptive and aversive dimensions of paclitaxel-induced peripheral neuropathy in female rats. Pain 2017, 158, 962–972. [Google Scholar] [CrossRef] [PubMed]
  65. Hamity, M.V.; Kolker, S.J.; Hegarty, D.M.; Blum, C.; Langmack, L.; Aicher, S.A.; Hammond, D.L. Nicotinamide riboside alleviates corneal and somatic hypersensitivity induced by paclitaxel in male rats. Invest. Ophthalmol. Vis. Sci. 2022, 63, 38. [Google Scholar] [CrossRef]
  66. Pîrvu, A.S.; Andrei, A.M.; Stănciulescu, E.C.; Baniţă, I.M.; Pisoschi, C.G.; Jurja, S.; Ciuluvica, R. NAD+ metabolism and retinal degeneration (Review). Exp. Ther. Med. 2021, 22, 670. [Google Scholar] [CrossRef]
  67. Zhang, X.; Zhang, N.; Chrenek, M.A.; Girardot, P.E.; Wang, J.; Sellers, J.T.; Geisert, E.E.; Brenner, C.; Nickerson, J.M.; Boatright, J.H.; et al. Systemic treatment with nicotinamide riboside is protective in two mouse models of retinal ganglion cell damage. Pharmaceutics 2021, 13, 893. [Google Scholar] [CrossRef]
  68. Zhang, X.; Henneman, N.F.; Girardot, P.E.; Sellers, J.T.; Chrenek, M.A.; Li, Y.; Wang, J.; Brenner, C.; Nickerson, J.M.; Boatright, J.H. Systemic treatment with nicotinamide riboside is protective in a mouse model of light-induced retinal degeneration. Invest. Ophthalmol. Vis. Sci. 2020, 61, 47. [Google Scholar] [CrossRef]
  69. Leung, C.K.S.; Ren, S.T.; Chan, P.P.M.; Wan, K.H.N.; Kam, A.K.W.; Lai, G.W.K.; Chiu, V.S.M.; Ko, M.W.L.; You, C.K.F.; Yu, M.C.Y. Nicotinamide riboside as a neuroprotective therapy for glaucoma: Study protocol for a randomized, double-blind, placebo-control trial. Trials 2022, 23, 45. [Google Scholar] [CrossRef]
  70. Silva, S.G.N.E.; Occhiutto, M.L.; Costa, V.P. The use of nicotinamide and nicotinamide riboside as an adjunct therapy in the treatment of glaucoma. Eur. J. Ophthalmol. 2023, 2023, 11206721231161101. [Google Scholar] [CrossRef]
  71. Zhou, B.; Zhao, G.; Zhu, Y.; Chen, X.; Zhang, N.; Yang, J.; Lin, H. Protective effects of nicotinamide riboside on H2O2-induced oxidative damage in lens epithelial cells. Curr. Eye Res. 2021, 46, 961–970. [Google Scholar] [CrossRef]
  72. Pang, H.; Jiang, Y.; Li, J.; Wang, Y.; Nie, M.; Xiao, N.; Wang, S.; Song, Z.; Ji, F.; Chang, Y.; et al. Aberrant NAD+ metabolism underlies Zika virus-induced microcephaly. Nat. Metab. 2021, 3, 1109–1124. [Google Scholar] [CrossRef]
  73. Matasic, D.S.; Brenner, C.; London, B. Emerging potential benefits of modulating NAD+ metabolism in cardiovascular disease. Am. J. Physiol. Heart Circ. Physiol. 2018, 314, H839–H852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Ahmad, F.; Tomar, D.; Aryal, A.C.S.; Elmoselhi, A.B.; Thomas, M.; Elrod, J.W.; Tilley, D.G.; Force, T. Nicotinamide riboside kinase-2 alleviates ischemia-induced heart failure through P38 signaling. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165609. [Google Scholar] [CrossRef] [PubMed]
  75. Blanco-Vaca, F.; Rotllan, N.; Canyelles, M.; Mauricio, D.; Escolà-Gil, J.C.; Julve, J. NAD+-increasing strategies to improve cardiometabolic health? Front. Endocrinol. 2022, 12, 815565. [Google Scholar] [CrossRef]
  76. Liu, Y.; Huang, Y.; Xu, C.; An, P.; Luo, Y.; Jiao, L.; Luo, J.; Li, Y. Mitochondrial dysfunction and therapeutic perspectives in cardiovascular diseases. Int. J. Mol. Sci. 2022, 23, 16053. [Google Scholar] [CrossRef] [PubMed]
  77. De Castro, J.M.; Assumpção, J.A.F.; Stein, D.J.; Toledo, R.S.; da Silva, L.S.; Caumo, W.; Carraro, C.C.; da Rosa Araujo, A.S.; Torres, I.L.S. Nicotinamide riboside reduces cardiometabolic risk factors and modulates cardiac oxidative stress in obese Wistar rats under caloric restriction. Life Sci. 2020, 263, 118596. [Google Scholar] [CrossRef]
  78. Liu, X.; Zhang, Y.; Deng, Y.; Yang, L.; Ou, W.; Xie, M.; Ding, L.; Jiang, C.; Yu, H.; Li, Q.; et al. Mitochondrial protein hyperacetylation underpins heart failure with preserved ejection fraction in mice. J. Mol. Cell Cardiol. 2022, 165, 76–85. [Google Scholar] [CrossRef]
  79. Ma, S.; Feng, J.; Lin, X.; Liu, J.; Tang, Y.; Nie, S.; Gong, J.; Wang, L. Nicotinamide riboside alleviates cardiac dysfunction and remodeling in pressure overload cardiac hypertrophy. Oxid. Med. Cell. Longev. 2021, 2021, 5546867. [Google Scholar] [CrossRef]
  80. Zhang, M.; Weng, H.; Zheng, J. NAD+ repletion inhibits the endothelial-to-mesenchymal transition induced by TGF-β in endothelial cells through improving mitochondrial unfolded protein response. Int. J. Biochem. Cell Biol. 2019, 117, 105635. [Google Scholar] [CrossRef]
  81. Pool, L.; Knops, P.; Manintveld, O.C.; Brugts, J.J.; Theuns, D.A.M.J.; Brundel, B.J.J.M.; de Groot, N.M.S. The HF-AF ENERGY Trial: Nicotinamide riboside for the treatment of atrial fibrillation in heart failure patients. Cardiovasc. Drugs Ther. 2022, 1–6. [Google Scholar] [CrossRef]
  82. Zhou, B.; Wang, D.D.; Qiu, Y.; Airhart, S.; Liu, Y.; Stempien-Otero, A.; O’Brien, K.D.; Tian, R. Boosting NAD level suppresses inflammatory activation of PBMCs in heart failure. J. Clin. Invest. 2020, 130, 6054–6063. [Google Scholar] [CrossRef]
  83. Wang, D.D.; Airhart, S.E.; Zhou, B.; Shireman, L.M.; Jiang, S.; Melendez Rodriguez, C.; Kirkpatrick, J.N.; Shen, D.D.; Tian, R.; O’Brien, K.D. Safety and tolerability of nicotinamide riboside in heart failure with reduced ejection fraction. JACC Basic Transl. Sci. 2022, 7, 1183–1196. [Google Scholar] [CrossRef]
  84. Oller, J.; Gabandé-Rodríguez, E.; Roldan-Montero, R.; Ruiz-Rodríguez, M.J.; Redondo, J.M.; Martín-Ventura, J.L.; Mittelbrunn, M. Rewiring vascular metabolism prevents sudden death due to aortic ruptures—Brief report. Arterioscler. Thromb. Vasc. Biol. 2022, 42, 462–469. [Google Scholar] [CrossRef] [PubMed]
  85. Shahzadi, S.K.; Marzook, H.; Qaisar, R.; Ahmad, F. Nicotinamide riboside kinase-2 inhibits JNK pathway and limits dilated cardiomyopathy in mice with chronic pressure overload. Clin. Sci. 2022, 136, 181–196. [Google Scholar] [CrossRef] [PubMed]
  86. Diguet, N.; Trammell, S.A.J.; Tannous, C.; Deloux, R.; Piquereau, J.; Mougenot, N.; Gouge, A.; Gressette, M.; Manoury, B.; Blanc, J.; et al. Nicotinamide riboside preserves cardiac function in a mouse model of dilated cardiomyopathy. Circulation 2018, 137, 2256–2273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Zheng, D.; Zhang, Y.; Zheng, M.; Cao, T.; Wang, G.; Zhang, L.; Ni, R.; Brockman, J.; Zhong, H.; Fan, G.C.; et al. Nicotinamide riboside promotes autolysosome clearance in preventing doxorubicin-induced cardiotoxicity. Clin. Sci. 2019, 133, 1505–1521. [Google Scholar] [CrossRef] [PubMed]
  88. Podyacheva, E.; Toropova, Y. Nicotinamide riboside for the prevention and treatment of doxorubicin cardiomyopathy. Opportunities and prospects. Nutrients 2021, 13, 3435. [Google Scholar] [CrossRef]
  89. Podyacheva, E.; Semenova, N.Y.; Zinserling, V.A.; Mukhametdinova, D.; Goncharova, I.; Zelinskaya, I.; Sviridov, E.; Martynov, M.; Osipova, S.; Toropova, Y. Intravenous nicotinamide riboside administration has a cardioprotective effect in chronic doxorubicin-induced cardiomyopathy. Int. J. Mol. Sci. 2022, 23, 13096. [Google Scholar] [CrossRef]
  90. Marzook, H.; Gupta, A.; Tomar, D.; Saleh, M.A.; Patil, K.; Semreen, M.H.; Hamoudi, R.; Soares, N.C.; Qaisar, R.; Ahmad, F. Nicotinamide riboside kinase-2 regulates metabolic adaptation in the ischemic heart. J. Mol. Med. 2023, 101, 311–326. [Google Scholar] [CrossRef]
  91. Nie, H.; Zhang, Y.; Yu, H.; Xiao, H.; Li, T.; Yang, Q. Oral delivery of carrier-free dual-drug nanocrystal self-assembled microspheres improved NAD+ bioavailability and attenuated cardiac ischemia/reperfusion injury in mice. Drug Deliv. 2021, 28, 433–444. [Google Scholar] [CrossRef]
  92. Xiao, Y.; Phelp, P.; Wang, Q.; Bakker, D.; Nederlof, R.; Hollmann, M.W.; Zuurbier, C.J. Cardioprotective properties of known agents in rat ischemia-reperfusion model under clinically relevant conditions: Only the NAD precursor nicotinamide riboside reduces infarct size in presence of fentanyl, midazolam and cangrelor, but not propofol. Front. Cardiovasc. Med. 2021, 8, 712478. [Google Scholar] [CrossRef] [PubMed]
  93. Toropova, Y.G.; Pechnikova, N.A.; Zelinskaya, I.A.; Zhuravsky, S.G.; Kornyushin, O.V.; Gonchar, A.I.; Ivkin, D.Y.; Leonova, Y.V.; Karev, V.E.; Karabak, I.A. Nicotinamide riboside has protective effects in a rat model of mesenteric ischaemia–reperfusion. Int. J. Exp. Pathol. 2018, 99, 304–311. [Google Scholar] [CrossRef]
  94. Mateuszuk, Ł.; Campagna, R.; Kutryb-Zając, B.; Kuś, K.; Słominska, E.M.; Smolenski, R.T.; Chlopicki, S. Reversal of endothelial dysfunction by nicotinamide mononucleotide via extracellular conversion to nicotinamide riboside. Biochem. Pharmacol. 2020, 178, 114019. [Google Scholar] [CrossRef] [PubMed]
  95. Mukherjee, S.; Chellappa, K.; Moffitt, A.; Ndungu, J.; Dellinger, R.W.; Davis, J.G.; Agarwal, B.; Baur, J.A. Nicotinamide adenine dinucleotide biosynthesis promotes liver regeneration. Hepatology 2017, 65, 616–630. [Google Scholar] [CrossRef] [PubMed]
  96. Dall, M.; Trammell, S.A.J.; Asping, M.; Hassing, A.S.; Agerholm, M.; Vienberg, S.G.; Gillum, M.P.; Larsen, S.; Treebak, J.T. Mitochondrial function in liver cells is resistant to perturbations in NAD+ salvage capacity. J. Biol. Chem. 2019, 294, 13304–13326. [Google Scholar] [CrossRef] [PubMed]
  97. Lee, H.J.; Yang, S.J. Nicotinamide riboside regulates inflammation and mitochondrial markers in AML12 hepatocytes. Nutr. Res. Pr. 2019, 13, 3–10. [Google Scholar] [CrossRef]
  98. Fan, R.; Cui, J.; Ren, F.; Wang, Q.; Huang, Y.; Zhao, B.; Wei, L.; Qian, X.; Xiong, X. Overexpression of NRK1 ameliorates diet- and age-induced hepatic steatosis and insulin resistance. Biochem. Biophys. Res. Commun. 2018, 500, 476–483. [Google Scholar] [CrossRef]
  99. Han, X.; Bao, X.; Lou, Q.; Xie, X.; Zhang, M.; Zhou, S.; Guo, H.; Jiang, G.; Shi, Q. Nicotinamide riboside exerts protective effect against aging-induced NAFLD-like hepatic dysfunction in mice. PeerJ 2019, 7, e7568. [Google Scholar] [CrossRef] [Green Version]
  100. Bock, K.W. Modulation of aryl hydrocarbon receptor (AHR) and the NAD+-consuming enzyme CD38: Searches of therapeutic options for nonalcoholic fatty liver disease (NAFLD). Biochem. Pharmacol. 2020, 175, 113905. [Google Scholar] [CrossRef]
  101. Dall, M.; Hassing, A.S.; Niu, L.; Nielsen, T.S.; Ingerslev, L.R.; Sulek, K.; Trammell, S.A.J.; Gillum, M.P.; Barrès, R.; Larsen, S.; et al. Hepatocyte-specific perturbation of NAD+ biosynthetic pathways in mice induces reversible nonalcoholic steatohepatitis-like phenotypes. J. Biol. Chem. 2021, 297, 101388. [Google Scholar] [CrossRef]
  102. Quesada-Vázquez, S.; Bone, C.; Saha, S.; Triguero, I.; Colom-Pellicer, M.; Aragonès, G.; Hildebrand, F.; Del Bas, J.M.; Caimari, A.; Beraza, N.; et al. Microbiota dysbiosis and gut barrier dysfunction associated with non-alcoholic fatty liver disease are modulated by a specific metabolic cofactors’ combination. Int. J. Mol. Sci. 2022, 23, 13675. [Google Scholar] [CrossRef]
  103. Dellinger, R.W.; Holmes, H.E.; Hu-Seliger, T.; Butt, R.W.; Harrison, S.A.; Mozaffarian, D.; Chen, O.; Guarente, L. Nicotinamide riboside and pterostilbene reduces markers of hepatic inflammation in NAFLD: A double-blind, placebo-controlled clinical trial. Hepatology 2022. [Google Scholar] [CrossRef] [PubMed]
  104. Pham, T.X.; Bae, M.; Kim, M.B.; Lee, Y.; Hu, S.; Kang, H.; Park, Y.K.; Lee, J.Y. Nicotinamide riboside, an NAD+ precursor, attenuates the development of liver fibrosis in a diet-induced mouse model of liver fibrosis. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 2451–2463. [Google Scholar] [CrossRef] [PubMed]
  105. Jiang, R.; Zhou, Y.; Wang, S.; Pang, N.; Huang, Y.; Ye, M.; Wan, T.; Qiu, Y.; Pei, L.; Jiang, X.; et al. Nicotinamide riboside protects against liver fibrosis induced by CCl4 via regulating the acetylation of Smads signaling pathway. Life Sci. 2019, 225, 20–28. [Google Scholar] [CrossRef]
  106. Wang, S.; Wan, T.; Ye, M.; Qiu, Y.; Pei, L.; Jiang, R.; Pang, N.; Huang, Y.; Liang, B.; Ling, W.; et al. Nicotinamide riboside attenuates alcohol induced liver injuries via activation of SirT1/PGC-1α/mitochondrial biosynthesis pathway. Redox Biol. 2018, 17, 89–98. [Google Scholar] [CrossRef]
  107. Yau, W.W.; Chen, G.B.; Zhou, J.; Francisco, J.C.; Thimmukonda, N.K.; Li, S.; Singh, B.K.; Yen, P.M. Nicotinamide riboside rescues dysregulated glycolysis and fatty acid β-oxidation in a human hepatic cell model of citrin deficiency. Human Mol. Genet. 2023, 32, 1922–1931. [Google Scholar] [CrossRef] [PubMed]
  108. Lee, S.R.; Roh, J.Y.; Ryu, J.; Shin, H.J.; Hong, E.J. Activation of TCA cycle restrains virus-metabolic hijacking and viral replication in mouse hepatitis virus-infected cells. Cell Biosci. 2022, 12, 7. [Google Scholar] [CrossRef]
  109. Parikh, S.M. Metabolic stress resistance in acute kidney injury: Evidence for a PPAR-gamma-coactivator-1 alpha-nicotinamide adenine dinucleotide pathway. Nephron 2019, 143, 184–187. [Google Scholar] [CrossRef]
  110. Simic, P.; Vela Parada, X.F.; Parikh, S.M.; Dellinger, R.; Guarente, L.P.; Rhee, E.P. Nicotinamide riboside with pterostilbene (NRPT) increases NAD+ in patients with acute kidney injury (AKI): A randomized, double-blind, placebo-controlled, stepwise safety study of escalating doses of NRPT in patients with AKI. BMC Nephrol. 2020, 21, 342. [Google Scholar] [CrossRef]
  111. Lin, W.; Wu, X.; Wen, J.; Fei, Y.; Wu, J.; Li, X.; Zhang, Q.; Dong, Y.; Xu, T.; Fan, Y.; et al. Nicotinamide retains Klotho expression and ameliorates rhabdomyolysis-induced acute kidney injury. Nutrition 2021, 91–92, 111376. [Google Scholar] [CrossRef]
  112. Piedrafita, A.; Balayssac, S.; Mayeur, N.; Gazut, S.; Grossac, J.; Buleon, M.; Alves, M.; Klein, J.; Minville, V.; Marcheix, B.; et al. The tryptophan pathway and nicotinamide supplementation in ischaemic acute kidney injury. Clin. Kidney J. 2021, 14, 2490–2496. [Google Scholar] [CrossRef] [PubMed]
  113. Fontecha-Barriuso, M.; Lopez-Diaz, A.M.; Carriazo, S.; Ortiz, A.; Sanz, A.B. Nicotinamide and acute kidney injury. Clin. Kidney J. 2021, 14, 2453–2462. [Google Scholar] [CrossRef] [PubMed]
  114. Pezzotta, A.; Perico, L.; Morigi, M.; Corna, D.; Locatelli, M.; Zoja, C.; Benigni, A.; Remuzzi, G.; Imberti, B. Low nephron number induced by maternal protein restriction is prevented by nicotinamide riboside supplementation depending on sirtuin 3 activation. Cells 2022, 11, 3316. [Google Scholar] [CrossRef] [PubMed]
  115. Morevati, M.; Egstrand, S.; Nordholm, A.; Mace, M.L.; Andersen, C.B.; Salmani, R.; Olgaard, K.; Lewin, E. Effect of NAD+ boosting on kidney ischemia–reperfusion injury. PLoS ONE 2021, 16, e0252554. [Google Scholar] [CrossRef] [PubMed]
  116. Doke, T.; Mukherjee, S.; Mukhi, D.; Dhillon, P.; Abedini, A.; Davis, J.G.; Chellappa, K.; Chen, B.; Baur, J.A.; Susztak, K. NAD+ precursor supplementation prevents mtRNA/RIG-I-dependent inflammation during kidney injury. Nat. Metab. 2023, 5, 414–430. [Google Scholar] [CrossRef] [PubMed]
  117. Ahmadi, A.; Begue, G.; Valencia, A.P.; Norman, J.E.; Lidgard, B.; Bennett, B.J.; Van Doren, M.P.; Marcinek, D.J.; Fan, S.; Prince, D.K.; et al. Randomized crossover clinical trial of coenzyme Q10 and nicotinamide riboside in chronic kidney disease. JCI Insight 2023, 8, e167274. [Google Scholar] [CrossRef]
  118. Fletcher, R.S.; Ratajczak, J.; Doig, C.L.; Oakey, L.A.; Callingham, R.; Da Silva Xavier, G.; Garten, A.; Elhassan, Y.S.; Redpath, P.; Migaud, M.E.; et al. Nicotinamide riboside kinases display redundancy in mediating nicotinamide mononucleotide and nicotinamide riboside metabolism in skeletal muscle cells. Mol. Metab. 2017, 6, 819–832. [Google Scholar] [CrossRef]
  119. Sonntag, T.; Ancel, S.; Karaz, S.; Cichosz, P.; Jacot, G.; Giner, M.P.; Sanchez-Garcia, J.L.; Pannérec, A.; Moco, S.; Sorrentino, V.; et al. Nicotinamide riboside kinases regulate skeletal muscle fiber-type specification and are rate-limiting for metabolic adaptations during regeneration. Front. Cell Dev. Biol. 2022, 10, 1049653. [Google Scholar] [CrossRef]
  120. Doig, C.L.; Zielinska, A.E.; Fletcher, R.S.; Oakey, L.A.; Elhassan, Y.S.; Garten, A.; Cartwright, D.; Heising, S.; Alsheri, A.; Watson, D.G.; et al. Induction of the nicotinamide riboside kinase NAD+ salvage pathway in a model of sarcoplasmic reticulum dysfunction. Skelet. Muscle 2020, 10, 5. [Google Scholar] [CrossRef] [Green Version]
  121. Damgaard, M.V.; Nielsen, T.S.; Basse, A.L.; Chubanava, S.; Trost, K.; Moritz, T.; Dellinger, R.W.; Larsen, S.; Treebak, J.T. Intravenous nicotinamide riboside elevates mouse skeletal muscle NAD+ without impacting respiratory capacity or insulin sensitivity. iScience 2022, 25, 103863. [Google Scholar] [CrossRef]
  122. Elhassan, Y.S.; Kluckova, K.; Fletcher, R.S.; Schmidt, M.S.; Garten, A.; Doig, C.L.; Cartwright, D.M.; Oakey, L.; Burley, C.V.; Jenkinson, N.; et al. Nicotinamide riboside augments the aged human skeletal muscle NAD+ metabolome and induces transcriptomic and anti-inflammatory signatures. Cell Rep. 2019, 28, 1717–1728.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Remie, C.M.E.; Roumans, K.H.M.; Moonen, M.P.B.; Connell, N.J.; Havekes, B.; Mevenkamp, J.; Lindeboom, L.; de Wit, V.H.W.; van de Weijer, T.; Aarts, S.A.B.M.; et al. Nicotinamide riboside supplementation alters body composition and skeletal muscle acetylcarnitine concentrations in healthy obese humans. Am. J. Clin. Nutr. 2020, 112, 413–426. [Google Scholar] [CrossRef] [PubMed]
  124. Custodero, C.; Saini, S.K.; Shin, M.J.; Jeon, Y.K.; Christou, D.D.; McDermott, M.M.; Leeuwenburgh, C.; Anton, S.D.; Mankowski, R.T. Nicotinamide riboside—A missing piece in the puzzle of exercise therapy for older adults? Exp. Gerontol. 2020, 137, 110972. [Google Scholar] [CrossRef]
  125. Campelj, D.; Philp, A. NAD+ therapeutics and skeletal muscle adaptation to exercise in humans. Sports Med. 2022, 52, 91–99. [Google Scholar] [CrossRef]
  126. Stocks, B.; Ashcroft, S.P.; Joanisse, S.; Dansereau, L.C.; Koay, Y.C.; Elhassan, Y.S.; Lavery, G.G.; Quek, L.E.; O’Sullivan, J.F.; Philp, A.M.; et al. Nicotinamide riboside supplementation does not alter whole-body or skeletal muscle metabolic responses to a single bout of endurance exercise. J. Physiol. 2021, 599, 1513–1531. [Google Scholar] [CrossRef] [PubMed]
  127. Song, M.; Armenian, S.H.; Bhandari, R.; Lee, K.; Ness, K.; Putt, M.; Lindenfeld, L.; Manoukian, S.; Wade, K.; Dedio, A.; et al. Exercise training and NR supplementation to improve muscle mass and fitness in adolescent and young adult hematopoietic cell transplant survivors: A randomized controlled trial {1}. BMC Cancer 2022, 22, 795. [Google Scholar] [CrossRef]
  128. Kourtzidis, I.A.; Stoupas, A.T.; Gioris, I.S.; Veskoukis, A.S.; Margaritelis, N.V.; Tsantarliotou, M.; Taitzoglou, I.; Vrabas, I.S.; Paschalis, V.; Kyparos, A.; et al. The NAD+ precursor nicotinamide riboside decreases exercise performance in rats. J. Int. Soc. Sports Nutr. 2016, 13, 32. [Google Scholar] [CrossRef] [Green Version]
  129. Kourtzidis, I.A.; Dolopikou, C.F.; Tsiftsis, A.N.; Margaritelis, N.V.; Theodorou, A.A.; Zervos, I.A.; Tsantarliotou, M.P.; Veskoukis, A.S.; Vrabas, I.S.; Paschalis, V.; et al. Nicotinamide riboside supplementation dysregulates redox and energy metabolism in rats: Implications for exercise performance. Exp. Physiol. 2018, 103, 1357–1366. [Google Scholar] [CrossRef] [Green Version]
  130. Deloux, R.; Tannous, C.; Ferry, A.; Li, Z.; Mericskay, M. Aged nicotinamide riboside kinase 2 deficient mice present an altered response to endurance exercise training. Front. Physiol. 2018, 9, 1290. [Google Scholar] [CrossRef] [Green Version]
  131. Zhang, H.; Ryu, D.; Wu, Y.; Gariani, K.; Wang, X.; Luan, P.; D’Amico, D.; Ropelle, E.R.; Lutolf, M.P.; Aebersold, R.; et al. NAD+ repletion improves mitochondrial and stem cell function and enhances life span in mice. Science 2016, 352, 1436–1443. [Google Scholar] [CrossRef] [Green Version]
  132. Moon, J.; Kim, H.R.; Shin, M.G. Rejuvenating aged hematopoietic stem cells through improvement of mitochondrial function. Ann. Lab. Med. 2018, 38, 395–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Khan, N.A.; Auranen, M.; Paetau, I.; Pirinen, E.; Euro, L.; Forsström, S.; Pasila, L.; Velagapudi, V.; Carroll, C.J.; Auwerx, J.; et al. Effective treatment of mitochondrial myopathy by nicotinamide riboside, a vitamin B3. EMBO Mol. Med. 2014, 6, 721–731. [Google Scholar] [CrossRef] [PubMed]
  134. Schaefer, P.M.; Huang, J.; Butic, A.; Perry, C.; Yardeni, T.; Tan, W.; Morrow, R.; Baur, J.A.; Wallace, D.C. Nicotinamide riboside alleviates exercise intolerance in ANT1-deficient mice. Mol. Metab. 2022, 64, 101560. [Google Scholar] [CrossRef] [PubMed]
  135. Lapatto, H.A.K.; Kuusela, M.; Heikkinen, A.; Muniandy, M.; van der Kolk, B.W.; Gopalakrishnan, S.; Pöllänen, N.; Sandvik, M.; Schmidt, M.S.; Heinonen, S.; et al. Nicotinamide riboside improves muscle mitochondrial biogenesis, satellite cell differentiation, and gut microbiota in a twin study. Sci. Adv. 2023, 9, eadd5163. [Google Scholar] [CrossRef]
  136. Perez-Sanchez, C.; Escudero-Contreras, A.; Cerdó, T.; Sánchez-Mendoza, L.M.; Llamas-Urbano, A.; Arias-de la Rosa, I.; Pérez-Rodriguez, M.; Muñoz-Barrera, L.; Abalos-Aguilera, M.D.C.; Barbarroja, N.; et al. Preclinical characterization of pharmacological NAD+ boosting as a promising therapeutic approach in rheumatoid arthritis. Arthritis Rheumatol. 2023. [Google Scholar] [CrossRef]
  137. Gonzalez, J.M.; Jackson, A.R. In ovo feeding of nicotinamide riboside affects broiler pectoralis major muscle development. Transl. Anim. Sci. 2020, 4, txaa126. [Google Scholar] [CrossRef]
  138. Xu, X.; Jackson, A.R.; Gonzalez, J.M. The effects of in ovo nicotinamide riboside dose on broiler myogenesis. Poult. Sci. 2021, 100, 100926. [Google Scholar] [CrossRef]
  139. Xu, X.; Alcocer, H.M.; Gravely, M.E.; Jackson, A.R.; Gonzalez, J.M. Effects of in ovo injection of nicotinamide riboside on high-yield broiler myogenesis. J. Anim. Sci. 2022, 100, skac203. [Google Scholar] [CrossRef]
  140. Sambeat, A.; Ratajczak, J.; Joffraud, M.; Sanchez-Garcia, J.L.; Giner, M.P.; Valsesia, A.; Giroud-Gerbetant, J.; Valera-Alberni, M.; Cercillieux, A.; Boutant, M.; et al. Endogenous nicotinamide riboside metabolism protects against diet-induced liver damage. Nat. Commun. 2019, 10, 4291. [Google Scholar] [CrossRef] [Green Version]
  141. Igarashi, M.; Miura, M.; Williams, E.; Jaksch, F.; Kadowaki, T.; Yamauchi, T.; Guarente, L. NAD+ supplementation rejuvenates aged gut adult stem cells. Aging Cell 2019, 18, e12935. [Google Scholar] [CrossRef] [Green Version]
  142. Cartwright, D.M.; Oakey, L.A.; Fletcher, R.S.; Doig, C.L.; Heising, S.; Larner, D.P.; Nasteska, D.; Berry, C.E.; Heaselgrave, S.R.; Ludwig, C.; et al. Nicotinamide riboside has minimal impact on energy metabolism in mouse models of mild obesity. J. Endocrinol. 2021, 251, 111–123. [Google Scholar] [CrossRef] [PubMed]
  143. Lozada-Fernández, V.V.; de Leon, O.; Kellogg, S.L.; Saravia, F.L.; Hadiono, M.A.; Atkinson, S.N.; Grobe, J.L.; Kirby, J.R. Nicotinamide riboside-conditioned microbiota deflects high-fat diet-induced weight gain in mice. mSystems 2022, 7, e0023021. [Google Scholar] [CrossRef] [PubMed]
  144. Sauve, A.A. Metabolic disease, NAD metabolism, nicotinamide riboside, and the gut microbiome: Connecting the dots from the gut to physiology. mSystems 2022, 7, e0122321. [Google Scholar] [CrossRef] [PubMed]
  145. Kolba, N.; Zarei, A.; Cheng, J.; Agarwal, N.; Dadmohammadi, Y.; Khazdooz, L.; Abbaspourrad, A.; Tako, E. Alterations in intestinal brush border membrane functionality and bacterial populations following intra-amniotic administration (Gallus gallus) of nicotinamide riboside and its derivatives. Nutrients 2022, 14, 3130. [Google Scholar] [CrossRef] [PubMed]
  146. Peluso, A.A.; Lundgaard, A.T.; Babaei, P.; Mousovich-Neto, F.; Rocha, A.L.; Damgaard, M.V.; Bak, E.G.; Gnanasekaran, T.; Dollerup, O.L.; Trammell, S.A.J.; et al. Oral supplementation of nicotinamide riboside alters intestinal microbial composition in rats and mice, but not humans. NPJ Aging 2023, 9, 7. [Google Scholar] [CrossRef]
  147. Kim, M.B.; Pham, T.X.; van Luling, M.; Kostour, V.; Kang, H.; Corvino, O.; Jang, H.; Odell, W.; Bae, M.; Park, Y.K.; et al. Nicotinamide riboside supplementation exerts an anti-obesity effect and prevents inflammation and fibrosis in white adipose tissue of female diet-induced obesity mice. J. Nutr. Biochem. 2022, 107, 109058. [Google Scholar] [CrossRef]
  148. Longo, L.; de Castro, J.M.; Keingeski, M.B.; Rampelotto, P.H.; Stein, D.J.; Guerreiro, G.T.S.; de Souza, V.E.G.; Cerski, C.T.S.; Uribe-Cruz, C.; Torres, I.L.S.; et al. Nicotinamide riboside and dietary restriction effects on gut microbiota and liver inflammatory and morphologic markers in cafeteria diet-induced obesity in rats. Nutrition 2023, 110, 112019. [Google Scholar] [CrossRef]
  149. Yu, X.; Xue, M.; Liu, Y.; Zhou, Z.; Jiang, Y.; Sun, T.; Liang, H. Effect of nicotinamide riboside on lipid metabolism and gut microflora–bile acid axis in alcohol-exposed mice. Food Sci. Nutr. 2020, 9, 429–440. [Google Scholar] [CrossRef]
  150. Crisol, B.M.; Veiga, C.B.; Lenhare, L.; Braga, R.R.; Silva, V.R.R.; da Silva, A.S.R.; Cintra, D.E.; Moura, L.P.; Pauli, J.R.; Ropelle, E.R. Nicotinamide riboside induces a thermogenic response in lean mice. Life Sci. 2018, 211, 1–7. [Google Scholar] [CrossRef]
  151. Crisol, B.M.; Veiga, C.B.; Braga, R.R.; Lenhare, L.; Baptista, I.L.; Gaspar, R.C.; Muñoz, V.R.; Cordeiro, A.V.; da Silva, A.S.R.; Cintra, D.E.; et al. NAD+ precursor increases aerobic performance in mice. Eur. J. Nutr. 2020, 59, 2427–2437. [Google Scholar] [CrossRef]
  152. Williams, A.S.; Koves, T.R.; Pettway, Y.D.; Draper, J.A.; Slentz, D.H.; Grimsrud, P.A.; Ilkayeva, O.R.; Muoio, D.M. Nicotinamide riboside supplementation confers marginal metabolic benefits in obese mice without remodeling the muscle acetyl-proteome. iScience 2021, 25, 103635. [Google Scholar] [CrossRef]
  153. Serrano, A.; Palou, A.; Bonet, M.L.; Ribot, J. Nicotinamide riboside supplementation to suckling male mice improves lipid and energy metabolism in skeletal muscle and liver in adulthood. Nutrients 2022, 14, 2259. [Google Scholar] [CrossRef]
  154. De Castro, J.M.; Stein, D.J.; Medeiros, H.R.; de Oliveira, C.; Torres, I.L.S. Nicotinamide riboside neutralizes hypothalamic inflammation and increases weight loss without altering muscle mass in obese rats under calorie restriction: A preliminary investigation. Front. Nutr. 2021, 8, 648893. [Google Scholar] [CrossRef] [PubMed]
  155. Shi, W.; Hegeman, M.A.; van Dartel, D.A.M.; Tang, J.; Suarez, M.; Swarts, H.; van der Hee, B.; Arola, L.; Keijer, J. Effects of a wide range of dietary nicotinamide riboside (NR) concentrations on metabolic flexibility and white adipose tissue (WAT) of mice fed a mildly obesogenic diet. Mol. Nutr. Food Res. 2017, 61, 1600878. [Google Scholar] [CrossRef] [Green Version]
  156. Shi, W.; Hegeman, M.A.; Doncheva, A.; Bekkenkamp-Grovenstein, M.; de Boer, V.C.J.; Keijer, J. High dose of dietary nicotinamide riboside induces glucose intolerance and white adipose tissue dysfunction in mice fed a mildly obesogenic diet. Nutrients 2019, 11, 2439. [Google Scholar] [CrossRef] [Green Version]
  157. Shi, W.; Hegeman, M.A.; Doncheva, A.; van der Stelt, I.; Bekkenkamp-Grovenstein, M.; van Schothorst, E.M.; Brenner, C.; de Boer, V.C.J.; Keijer, J. Transcriptional response of white adipose tissue to withdrawal of vitamin B3. Mol. Nutr. Food Res. 2019, 63, e1801100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Quesada-Vázquez, S.; Antolín, A.; Colom-Pellicer, M.; Aragonès, G.; Herrero, L.; Del Bas, J.M.; Caimari, A.; Escoté, X. Reduction of obesity and insulin resistance through dual targeting of VAT and BAT by a novel combination of metabolic cofactors. Int. J. Mol. Sci. 2022, 23, 14923. [Google Scholar] [CrossRef]
  159. Zhao, H.; Tian, Y.; Zuo, Y.; Zhang, X.; Gao, Y.; Wang, P.; Sun, L.; Zhang, H.; Liang, H. Nicotinamide riboside ameliorates high-fructose-induced lipid metabolism disorder in mice via improving FGF21 resistance in the liver and white adipose tissue. Food Funct. 2022, 13, 12400–12411. [Google Scholar] [CrossRef] [PubMed]
  160. Serrano, A.; Asnani-Kishnani, M.; Couturier, C.; Astier, J.; Palou, A.; Landrier, J.F.; Ribot, J.; Bonet, M.L. DNA methylation changes are associated with the programming of white adipose tissue browning features by resveratrol and nicotinamide riboside neonatal supplementations in mice. Nutrients 2020, 12, 461. [Google Scholar] [CrossRef] [Green Version]
  161. Salic, K.; Gart, E.; Seidel, F.; Verschuren, L.; Caspers, M.; van Duyvenvoorde, W.; Wong, K.E.; Keijer, J.; Bobeldijk-Pastorova, I.; Wielinga, P.Y.; et al. Combined treatment with L-carnitine and nicotinamide riboside improves hepatic metabolism and attenuates obesity and liver steatosis. Int. J. Mol. Sci. 2019, 20, 4359. [Google Scholar] [CrossRef] [Green Version]
  162. Nagy, L.; Rauch, B.; Szerafin, T.; Uray, K.; Tóth, A.; Bai, P. Nicotinamide-riboside shifts the differentiation of human primary white adipocytes to beige adipocytes impacting substrate preference and uncoupling respiration through SIRT1 activation and mitochondria-derived reactive species production. Front. Cell Dev. Biol. 2022, 10, 979330. [Google Scholar] [CrossRef]
  163. Nascimento, E.B.M.; Moonen, M.P.B.; Remie, C.M.E.; Gariani, K.; Jörgensen, J.A.; Schaart, G.; Hoeks, J.; Auwerx, J.; van Marken Lichtenbelt, W.D.; Schrauwen, P. Nicotinamide riboside enhances in vitro beta-adrenergic brown adipose tissue activity in humans. J. Clin. Endocrinol. Metab. 2021, 106, 1437–1447. [Google Scholar] [CrossRef] [PubMed]
  164. Dollerup, O.L.; Trammell, S.A.J.; Hartmann, B.; Holst, J.J.; Christensen, B.; Møller, N.; Gillum, M.P.; Treebak, J.T.; Jessen, N. Effects of nicotinamide riboside on endocrine pancreatic function and incretin hormones in nondiabetic men with obesity. J. Clin. Endocrinol. Metab. 2019, 104, 5703–5714. [Google Scholar] [CrossRef] [PubMed]
  165. Dollerup, O.L.; Chubanava, S.; Agerholm, M.; Søndergård, S.D.; Altıntaş, A.; Møller, A.B.; Høyer, K.F.; Ringgaard, S.; Stødkilde-Jørgensen, H.; Lavery, G.G.; et al. Nicotinamide riboside does not alter mitochondrial respiration, content or morphology in skeletal muscle from obese and insulin-resistant men. J. Physiol. 2020, 598, 731–754. [Google Scholar] [CrossRef]
  166. Trammell, S.A.J.; Weidemann, B.J.; Chadda, A.; Yorek, M.S.; Holmes, A.; Coppey, L.J.; Obrosov, A.; Kardon, R.H.; Yorek, M.A.; Brenner, C. Nicotinamide riboside opposes type 2 diabetes and neuropathy in mice. Sci. Rep. 2016, 6, 26933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Chandrasekaran, K.; Najimi, N.; Sagi, A.R.; Yarlagadda, S.; Salimian, M.; Arvas, M.I.; Hedayat, A.F.; Kevas, Y.; Kadakia, A.; Russell, J.W. NAD+ precursors repair mitochondrial function in diabetes and prevent experimental diabetic neuropathy. Int. J. Mol. Sci. 2022, 23, 4887. [Google Scholar] [CrossRef]
  168. Cercillieux, A.; Ratajczak, J.; Joffraud, M.; Sanchez-Garcia, J.L.; Jacot, G.; Zollinger, A.; Métairon, S.; Giroud-Gerbetant, J.; Rumpler, M.; Ciarlo, E.; et al. Nicotinamide riboside kinase 1 protects against diet and age-induced pancreatic β-cell failure. Mol. Metab. 2022, 66, 101605. [Google Scholar] [CrossRef]
  169. Hu, L.; Guo, Y.; Song, L.; Wen, H.; Sun, N.; Wang, Y.; Qi, B.; Liang, Q.; Geng, J.; Liu, X.; et al. Nicotinamide riboside promotes Mfn2-mediated mitochondrial fusion in diabetic hearts through the SIRT1–PGC1α–PPARα pathway. Free Radic. Biol. Med. 2022, 183, 75–88. [Google Scholar] [CrossRef]
  170. Costa, C.J.; Cohen, M.W.; Goldberg, D.C.; Mellado, W.; Willis, D.E. Nicotinamide riboside improves enteric neuropathy in streptozocin-induced diabetic rats through myenteric plexus neuroprotection. Dig. Dis. Sci. 2023, 68, 2963–2974. [Google Scholar] [CrossRef]
  171. Lee, H.J.; Hong, Y.S.; Jun, W.; Yang, S.J. Nicotinamide riboside ameliorates hepatic metaflammation by modulating NLRP3 inflammasome in a rodent model of type 2 diabetes. J. Med. Food 2015, 18, 1207–1213. [Google Scholar] [CrossRef]
  172. Wang, Z.H.; Bao, X.G.; Hu, J.J.; Shen, S.B.; Xu, G.H.; Wu, Y.L. Nicotinamide riboside enhances endothelial precursor cell function to promote refractory wound healing through mediating the Sirt1/AMPK pathway. Front. Pharmacol. 2021, 12, 671563. [Google Scholar] [CrossRef] [PubMed]
  173. Lee, S.R.; Jeong, S.H.; Mukae, M.; Kim, S.Y.; Ko, J.W.; Kwun, H.J.; Hong, E.J. Dietary supplementation with nicotinamide riboside improves fetal growth under hypoglycemia. J. Nutr. Biochem. 2023, 116, 109310. [Google Scholar] [CrossRef] [PubMed]
  174. Chu, X.; Raju, R.P. Regulation of NAD+ metabolism in aging and disease. Metabolism 2022, 126, 154923. [Google Scholar] [CrossRef] [PubMed]
  175. Ji, L.L.; Yeo, D. Maintenance of NAD+ homeostasis in skeletal muscle during aging and exercise. Cells 2022, 11, 710. [Google Scholar] [CrossRef]
  176. Gomes, A.P.; Price, N.L.; Ling, A.J.Y.; Moslehi, J.J.; Montgomery, M.K.; Rajman, L.; White, J.P.; Teodoro, J.S.; Wrann, C.D.; Hubbard, B.P.; et al. Declining NAD+ induces a pseudohypoxic state disrupting nuclear–mitochondrial communication during aging. Cell 2013, 155, 1624–1638. [Google Scholar] [CrossRef] [Green Version]
  177. Cercillieux, A.; Ciarlo, E.; Canto, C. Balancing NAD+ deficits with nicotinamide riboside: Therapeutic possibilities and limitations. Cell Mol. Life Sci. 2022, 79, 463. [Google Scholar] [CrossRef]
  178. Li, W.; Zhou, Y.; Pang, N.; Hu, Q.; Li, Q.; Sun, Y.; Ding, Y.; Gu, Y.; Xiao, Y.; Gao, M.; et al. NAD supplement alleviates intestinal barrier injury induced by ethanol via protecting epithelial mitochondrial function. Nutrients 2022, 15, 174. [Google Scholar] [CrossRef]
  179. Vannini, N.; Campos, V.; Girotra, M.; Trachsel, V.; Rojas-Sutterlin, S.; Tratwal, J.; Ragusa, S.; Stefanidis, E.; Ryu, D.; Rainer, P.Y.; et al. The NAD-booster nicotinamide riboside potently stimulates hematopoiesis through increased mitochondrial clearance. Cell Stem Cell 2019, 24, 405–418.e7. [Google Scholar] [CrossRef] [Green Version]
  180. Zong, L.; Tanaka-Yano, M.; Park, B.; Yanai, H.; Turhan, F.T.; Croteau, D.L.; Tian, J.; Fang, E.F.; Bohr, V.A.; Beerman, I. NAD+ augmentation with nicotinamide riboside improves lymphoid potential of Atm−/− and old mice HSCs. NPJ Aging Mech. Dis. 2021, 7, 25. [Google Scholar] [CrossRef]
  181. Sun, X.; Cao, B.; Naval-Sanchez, M.; Pham, T.; Sun, Y.B.Y.; Williams, B.; Heazlewood, S.Y.; Deshpande, N.; Li, J.; Kraus, F.; et al. Nicotinamide riboside attenuates age-associated metabolic and functional changes in hematopoietic stem cells. Nat. Commun. 2021, 12, 2665. [Google Scholar] [CrossRef]
  182. Yoshino, J.; Mills, K.F.; Yoon, M.J.; Imai, S. Nicotinamide mononucleotide, a key NAD+ intermediate, treats the pathophysiology of diet- and age-induced diabetes in mice. Cell Metab. 2011, 14, 528–536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Zapata-Pérez, R.; Tammaro, A.; Schomakers, B.V.; Scantlebery, A.M.L.; Denis, S.; Elfrink, H.L.; Giroud-Gerbetant, J.; Cantó, C.; López-Leonardo, C.; McIntyre, R.L.; et al. Reduced nicotinamide mononucleotide is a new and potent NAD+ precursor in mammalian cells and mice. FASEB J. 2021, 35, e21456. [Google Scholar] [CrossRef] [PubMed]
  184. Freeberg, K.A.; Craighead, D.H.; Martens, C.R.; You, Z.; Chonchol, M.; Seals, D.R. Nicotinamide riboside supplementation for treating elevated systolic blood pressure and arterial stiffness in midlife and older adults. Front. Cardiovasc. Med. 2022, 9, 881703. [Google Scholar] [CrossRef] [PubMed]
  185. Acklin, S.; Sadhukhan, R.; Du, W.; Patra, M.; Cholia, R.; Xia, F. Nicotinamide riboside alleviates cisplatin-induced peripheral neuropathy via SIRT2 activation. Neurooncol. Adv. 2022, 4, vdac101. [Google Scholar] [CrossRef]
  186. Mukherjee, S.; Mo, J.; Paolella, L.M.; Perry, C.E.; Toth, J.; Hugo, M.M.; Chu, Q.; Tong, Q.; Chellappa, K.; Baur, J.A. SIRT3 is required for liver regeneration but not for the beneficial effect of nicotinamide riboside. JCI Insight 2021, 6, e147193. [Google Scholar] [CrossRef]
  187. Zhang, Q.; Liu, X.; Li, N.; Zhang, J.; Yang, J.; Bu, P. Sirtuin 3 deficiency aggravates contrast-induced acute kidney injury. J. Transl. Med. 2018, 16, 313. [Google Scholar] [CrossRef]
  188. Kang, H.; Park, Y.K.; Lee, J.Y. Nicotinamide riboside, an NAD+ precursor, attenuates inflammation and oxidative stress by activating sirtuin 1 in alcohol-stimulated macrophages. Lab. Invest. 2021, 101, 1225–1237. [Google Scholar] [CrossRef]
  189. Seldeen, K.L.; Shahini, A.; Thiyagarajan, R.; Redae, Y.; Leiker, M.; Rajabian, N.; Dynka, A.; Andreadis, S.T.; Troen, B.R. Short-term nicotinamide riboside treatment improves muscle quality and function in mice and increases cellular energetics and differentiating capacity of myogenic progenitors. Nutrition 2021, 87–88, 111189. [Google Scholar] [CrossRef]
  190. Jensen, J.B.; Dollerup, O.L.; Møller, A.B.; Billeskov, T.B.; Dalbram, E.; Chubanava, S.; Damgaard, M.V.; Dellinger, R.W.; Trošt, K.; Moritz, T.; et al. A randomized placebo-controlled trial of nicotinamide riboside and pterostilbene supplementation in experimental muscle injury in elderly individuals. JCI Insight 2022, 7, e158314. [Google Scholar] [CrossRef]
  191. Dolopikou, C.F.; Kourtzidis, I.A.; Margaritelis, N.V.; Vrabas, I.S.; Koidou, I.; Kyparos, A.; Theodorou, A.A.; Paschalis, V.; Nikolaidis, M.G. Acute nicotinamide riboside supplementation improves redox homeostasis and exercise performance in old individuals: A double-blind cross-over study. Eur. J. Nutr. 2020, 59, 505–515. [Google Scholar] [CrossRef]
  192. Grezella, C.; Fernandez-Rebollo, E.; Franzen, J.; Ventura Ferreira, M.S.; Beier, F.; Wagner, W. Effects of senolytic drugs on human mesenchymal stromal cells. Stem Cell Res. Ther. 2018, 9, 108. [Google Scholar] [CrossRef] [Green Version]
  193. Delabie, W.; Maes, W.; Devloo, R.; Van den Hauwe, M.R.; Vanhoorelbeke, K.; Compernolle, V.; Feys, H.B. The senotherapeutic nicotinamide riboside raises platelet nicotinamide adenine dinucleotide levels but cannot prevent storage lesion. Transfusion 2020, 60, 165–174. [Google Scholar] [CrossRef] [Green Version]
  194. Harrison, D.E.; Strong, R.; Reifsnyder, P.; Kumar, N.; Fernandez, E.; Flurkey, K.; Javors, M.A.; Lopez-Cruzan, M.; Macchiarini, F.; Nelson, J.F.; et al. 17-α-Estradiol late in life extends lifespan in aging UM-HET3 male mice; nicotinamide riboside and three other drugs do not affect lifespan in either sex. Aging Cell 2021, 20, e13328. [Google Scholar] [CrossRef]
  195. Belenky, P.A.; Moga, T.G.; Brenner, C. Saccharomyces cerevisiae YOR071C encodes the high affinity nicotinamide riboside transporter Nrt1. J. Biol. Chem. 2008, 283, 8075–8079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Orlandi, I.; Alberghina, L.; Vai, M. Nicotinamide, nicotinamide riboside and nicotinic acid-emerging roles in replicative and chronological aging in yeast. Biomolecules 2020, 10, 604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Odoh, C.K.; Guo, X.; Arnone, J.T.; Wang, X.; Zhao, Z.K. The role of NAD and NAD precursors on longevity and lifespan modulation in the budding yeast, Saccharomyces cerevisiae. Biogerontology 2022, 23, 169–199. [Google Scholar] [CrossRef] [PubMed]
  198. Belenky, P.; Christensen, K.C.; Gazzaniga, F.; Pletnev, A.A.; Brenner, C. Nicotinamide riboside and nicotinic acid riboside salvage in fungi and mammals. Quantitative basis for Urh1 and purine nucleoside phosphorylase function in NAD+ metabolism. J. Biol. Chem. 2009, 284, 158–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Braidy, N.; Poljak, A.; Grant, R.; Jayasena, T.; Mansour, H.; Chan-Ling, T.; Guillemin, G.J.; Smythe, G.; Sachdev, P. Mapping NAD+ metabolism in the brain of ageing Wistar rats: Potential targets for influencing brain senescence. Biogerontology 2014, 15, 177–198. [Google Scholar] [CrossRef]
  200. Verdin, E. NAD+ in aging, metabolism, and neurodegeneration. Science 2015, 350, 1208–1213. [Google Scholar] [CrossRef]
  201. Reiten, O.K.; Wilvang, M.A.; Mitchell, S.J.; Hu, Z.; Fang, E.F. Preclinical and clinical evidence of NAD+ precursors in health, disease, and ageing. Mech. Ageing Dev. 2021, 199, 111567. [Google Scholar] [CrossRef]
  202. Braidy, N.; Liu, Y. Can nicotinamide riboside protect against cognitive impairment? Curr. Opin. Clin. Nutr. Metab. Care 2020, 23, 413–420. [Google Scholar] [CrossRef]
  203. Birkisdóttir, M.B.; van Galen, I.; Brandt, R.M.C.; Barnhoorn, S.; van Vliet, N.; van Dijk, C.; Nagarajah, B.; Imholz, S.; van Oostrom, C.T.; Reiling, E.; et al. The use of progeroid DNA repair-deficient mice for assessing anti-aging compounds, illustrating the benefits of nicotinamide riboside. Front. Aging 2022, 3, 1005322. [Google Scholar] [CrossRef] [PubMed]
  204. Hou, Y.; Lautrup, S.; Cordonnier, S.; Wang, Y.; Croteau, D.L.; Zavala, E.; Zhang, Y.; Moritoh, K.; O’Connell, J.F.; Baptiste, B.A.; et al. NAD+ supplementation normalizes key Alzheimer’s features and DNA damage responses in a new AD mouse model with introduced DNA repair deficiency. Proc. Natl. Acad. Sci. USA 2018, 115, E1876–E1885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Xie, X.; Gao, Y.; Zeng, M.; Wang, Y.; Wei, T.F.; Lu, Y.B.; Zhang, W.P. Nicotinamide ribose ameliorates cognitive impairment of aged and Alzheimer’s disease model mice. Metab. Brain Dis. 2019, 34, 353–366. [Google Scholar] [CrossRef]
  206. Larrick, J.W.; Mendelsohn, A.R. Modulation of cGAS-STING pathway by nicotinamide riboside in Alzheimer’s disease. Rejuvenation Res. 2021, 24, 397–402. [Google Scholar] [CrossRef]
  207. Roboon, J.; Hattori, T.; Ishii, H.; Takarada-Iemata, M.; Nguyen, D.T.; Heer, C.D.; O’Meally, D.; Brenner, C.; Yamamoto, Y.; Okamoto, H.; et al. Inhibition of CD38 and supplementation of nicotinamide riboside ameliorate lipopolysaccharide-induced microglial and astrocytic neuroinflammation by increasing NAD. J. Neurochem. 2021, 158, 311–327. [Google Scholar] [CrossRef] [PubMed]
  208. Chu, X.; Hou, Y.; Meng, Q.; Croteau, D.L.; Wei, Y.; De, S.; Becker, K.G.; Bohr, V.A. Nicotinamide adenine dinucleotide supplementation drives gut microbiota variation in Alzheimer’s mouse model. Front. Aging Neurosci. 2022, 14, 993615. [Google Scholar] [CrossRef]
  209. Ryu, W.I.; Shen, M.; Lee, Y.; Healy, R.A.; Bormann, M.K.; Cohen, B.M.; Sonntag, K.C. Nicotinamide riboside and caffeine partially restore diminished NAD availability but not altered energy metabolism in Alzheimer’s disease. Aging Cell 2022, 21, e13658. [Google Scholar] [CrossRef]
  210. Yulug, B.; Altay, O.; Li, X.; Hanoglu, L.; Cankaya, S.; Lam, S.; Velioglu, H.A.; Yang, H.; Coskun, E.; Idil, E.; et al. Combined metabolic activators improve cognitive functions in Alzheimer’s disease patients: A randomised, double-blinded, placebo-controlled phase-II trial. Transl. Neurodegener. 2023, 12, 4. [Google Scholar] [CrossRef]
  211. Li, C.C.; Chen, W.X.; Wang, J.; Xia, M.; Jia, Z.C.; Guo, C.; Tang, X.Q.; Li, M.X.; Yin, Y.; Liu, X.; et al. Nicotinamide riboside rescues angiotensin II-induced cerebral small vessel disease in mice. CNS Neurosci. Ther. 2020, 26, 438–447. [Google Scholar] [CrossRef] [Green Version]
  212. Schöndorf, D.C.; Ivanyuk, D.; Baden, P.; Sanchez-Martinez, A.; De Cicco, S.; Yu, C.; Giunta, I.; Schwarz, L.K.; Di Napoli, G.; Panagiotakopoulou, V.; et al. The NAD+ precursor nicotinamide riboside rescues mitochondrial defects and neuronal loss in iPSC and fly models of Parkinson’s disease. Cell Rep. 2018, 23, 2976–2988. [Google Scholar] [CrossRef] [PubMed]
  213. Brakedal, B.; Dölle, C.; Riemer, F.; Ma, Y.; Nido, G.S.; Skeie, G.O.; Craven, A.R.; Schwarzlmüller, T.; Brekke, N.; Diab, J.; et al. The NADPARK study: A randomized phase I trial of nicotinamide riboside supplementation in Parkinson’s disease. Cell Metab. 2022, 34, 396–407.e6. [Google Scholar] [CrossRef] [PubMed]
  214. Vreones, M.; Mustapic, M.; Moaddel, R.; Pucha, K.A.; Lovett, J.; Seals, D.R.; Kapogiannis, D.; Martens, C.R. Oral nicotinamide riboside raises NAD+ and lowers biomarkers of neurodegenerative pathology in plasma extracellular vesicles enriched for neuronal origin. Aging Cell 2023, 22, e13754. [Google Scholar] [CrossRef]
  215. Gaare, J.J.; Dölle, C.; Brakedal, B.; Brügger, K.; Haugarvoll, K.; Nido, G.S.; Tzoulis, C. Nicotinamide riboside supplementation is not associated with altered methylation homeostasis in Parkinson’s disease. iScience 2023, 26, 106278. [Google Scholar] [CrossRef]
  216. Pang, N.; Hu, Q.; Zhou, Y.; Xiao, Y.; Li, W.; Ding, Y.; Chen, Y.; Ye, M.; Pei, L.; Li, Q.; et al. Nicotinamide adenine dinucleotide precursor suppresses hepatocellular cancer progression in mice. Nutrients 2023, 15, 1447. [Google Scholar] [CrossRef]
  217. Park, J.M.; Han, Y.M.; Lee, H.J.; Park, Y.J.; Hahm, K.B. Nicotinamide riboside vitamin B3 mitigated C26 adenocarcinoma-induced cancer cachexia. Front. Pharmacol. 2021, 12, 665493. [Google Scholar] [CrossRef]
  218. Hamity, M.V.; White, S.R.; Blum, C.; Gibson-Corley, K.N.; Hammond, D.L. Nicotinamide riboside relieves paclitaxel-induced peripheral neuropathy and enhances suppression of tumor growth in tumor-bearing rats. Pain 2020, 161, 2364–2375. [Google Scholar] [CrossRef] [PubMed]
  219. Kahn, B.; Borrelli, M.; Libby, T. A narrative review of nicotinamide adenine dinucleotide (NAD)+ intermediates nicotinamide riboside and nicotinamide mononucleotide for keratinocyte carcinoma risk reduction. J. Drugs Dermatol. 2022, 21, 1129–1132. [Google Scholar] [CrossRef] [PubMed]
  220. Conze, D.B.; Crespo-Barreto, J.; Kruger, C.L. Safety assessment of nicotinamide riboside, a form of vitamin B3. Human Exp. Toxicol. 2016, 35, 1149–1160. [Google Scholar] [CrossRef] [Green Version]
  221. Dellinger, R.W.; Santos, S.R.; Morris, M.; Evans, M.; Alminana, D.; Guarente, L.; Marcotulli, E. Repeat dose NRPT (nicotinamide riboside and pterostilbene) increases NAD+ levels in humans safely and sustainably: A randomized, double-blind, placebo-controlled study. NPJ Aging Mech. Dis. 2017, 3, 17. [Google Scholar] [CrossRef] [Green Version]
  222. EFSA Panel on Nutrition, Novel Foods and Food Allergens (NDA); Turck, D.; Castenmiller, J.; de Henauw, S.; Hirsch-Ernst, K.I.; Kearney, J.; Maciuk, A.; Mangelsdorf, I.; McArdle, H.J.; Naska, A.; et al. Safety of nicotinamide riboside chloride as a novel food pursuant to Regulation (EU) 2015/2283 and bioavailability of nicotinamide from this source, in the context of Directive 2002/46/EC. EFSA J. 2019, 17, e05775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Marinescu, A.G.; Chen, J.; Holmes, H.E.; Guarente, L.; Mendes, O.; Morris, M.; Dellinger, R.W. Safety assessment of high-purity, synthetic nicotinamide riboside (NR-E) in a 90-day repeated dose oral toxicity study, with a 28-day recovery arm. Int. J. Toxicol. 2020, 39, 307–320. [Google Scholar] [CrossRef] [PubMed]
  224. EFSA Panel on Nutrition, Novel Foods and Food Allergens (NDA); Turck, D.; Bohn, T.; Castenmiller, J.; De Henauw, S.; Hirsch-Ernst, K.I.; Maciuk, A.; Mangelsdorf, I.; McArdle, H.J.; Naska, A.; et al. Extension of use of nicotinamide riboside chloride as a novel food pursuant to Regulation (EU) 2015/2283. EFSA J. 2021, 19, e06843. [Google Scholar] [CrossRef]
  225. Sun, P.; Qie, S.; Pan, B. Nicotinamide riboside will play an important role in anti-aging therapy in humans, especially in the face skin anti-aging treatment. Aesthetic. Plast Surg. 2022, 46, 192–194. [Google Scholar] [CrossRef] [PubMed]
  226. Ito, T.K.; Sato, T.; Hakamata, A.; Onoda, Y.; Sato, S.; Yamazaki, F.; Horikawa, M.; Takahashi, Y.; Kitamoto, T.; Suzuki, M.; et al. A nonrandomized study of single oral supplementation within the daily tolerable upper level of nicotinamide affects blood nicotinamide and NAD+ levels in healthy subjects. Transl. Med. Aging 2020, 4, 45–54. [Google Scholar] [CrossRef]
  227. Ito, T.K.; Sato, T.; Takanashi, Y.; Tamannaa, Z.; Kitamoto, T.; Odagiri, K.; Setou, M. A single oral supplementation of nicotinamide within the daily tolerable upper level increases blood NAD+ levels in healthy subjects. Transl. Med. Aging 2021, 5, 43–51. [Google Scholar] [CrossRef]
  228. Turner, J.; Licollari, A.; Mihalcea, E.; Tan, A. Safety evaluation for Restorin® NMN, a NAD+ precursor. Front. Pharmacol. 2021, 12, 749727. [Google Scholar] [CrossRef]
  229. Okabe, K.; Yaku, K.; Uchida, Y.; Fukamizu, Y.; Sato, T.; Sakurai, T.; Tobe, K.; Nakagawa, T. Oral administration of nicotinamide mononucleotide is safe and efficiently increases blood nicotinamide adenine dinucleotide levels in healthy subjects. Front. Nutr. 2022, 9, 868640. [Google Scholar] [CrossRef]
  230. Pencina, K.; Lavu, S.; Dos Santos, M.; Beleva, Y.M.; Cheng, M.; Livingston, D.; Bhasin, S. MIB-626, an oral formulation of a microcrystalline unique polymorph of β-nicotinamide mononucleotide, increases circulating nicotinamide adenine dinucleotide and its metabolome in middle-aged and older adults. J. Gerontol. A Biol. Sci. Med. Sci. 2022, 78, 90–96. [Google Scholar] [CrossRef]
  231. Trammell, S.A.J.; Schmidt, M.S.; Weidemann, B.J.; Redpath, P.; Jaksch, F.; Dellinger, R.W.; Li, Z.; Abel, E.D.; Migaud, M.E.; Brenner, C. Nicotinamide riboside is uniquely and orally bioavailable in mice and humans. Nat. Commun. 2016, 7, 12948. [Google Scholar] [CrossRef] [Green Version]
  232. Airhart, S.E.; Shireman, L.M.; Risler, L.J.; Anderson, G.D.; Nagana Gowda, G.A.; Raftery, D.; Tian, R.; Shen, D.D.; O’Brien, K.D. An open-label, non-randomized study of the pharmacokinetics of the nutritional supplement nicotinamide riboside (NR) and its effects on blood NAD+ levels in healthy volunteers. PLoS ONE 2017, 12, e0186459. [Google Scholar] [CrossRef] [PubMed]
  233. Campbell, M.T.D.; Jones, D.S.; Andrews, G.P.; Li, S. Understanding the physicochemical properties and degradation kinetics of nicotinamide riboside, a promising vitamin B3 nutritional supplement. Food Nutr. Res. 2019, 63, 3419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Yaku, K.; Palikhe, S.; Izumi, H.; Yoshida, T.; Hikosaka, K.; Hayat, F.; Karim, M.; Iqbal, T.; Nitta, Y.; Sato, A.; et al. BST1 regulates nicotinamide riboside metabolism via its glycohydrolase and base-exchange activities. Nat. Commun. 2021, 12, 6767. [Google Scholar] [CrossRef] [PubMed]
  235. Chellappa, K.; McReynolds, M.R.; Lu, W.; Zeng, X.; Makarov, M.; Hayat, F.; Mukherjee, S.; Bhat, Y.R.; Lingala, S.R.; Shima, R.T.; et al. NAD precursors cycle between host tissues and the gut microbiome. Cell Metab. 2022, 34, 1947–1959.e5. [Google Scholar] [CrossRef]
  236. Preugschat, F.; Carter, L.H.; Boros, E.E.; Porter, D.J.; Stewart, E.L.; Shewchuk, L.M. A pre-steady state and steady state kinetic analysis of the N-ribosyl hydrolase activity of hCD157. Arch Biochem. Biophys. 2014, 564, 156–163. [Google Scholar] [CrossRef] [PubMed]
  237. Kropotov, A.; Kulikova, V.; Nerinovski, K.; Yakimov, A.; Svetlova, M.; Solovjeva, L.; Sudnitsyna, J.; Migaud, M.E.; Khodorkovskiy, M.; Ziegler, M.; et al. Equilibrative nucleoside transporters mediate the import of nicotinamide riboside and nicotinic acid riboside into human cells. Int. J. Mol. Sci. 2021, 22, 1391. [Google Scholar] [CrossRef]
  238. Kropotov, A.; Kulikova, V.; Solovjeva, L.; Yakimov, A.; Nerinovski, K.; Svetlova, M.; Sudnitsyna, J.; Plusnina, A.; Antipova, M.; Khodorkovskiy, M.; et al. Purine nucleoside phosphorylase controls nicotinamide riboside metabolism in mammalian cells. J. Biol. Chem. 2022, 298, 102615. [Google Scholar] [CrossRef]
  239. Kulikova, V.; Shabalin, K.; Nerinovski, K.; Dölle, C.; Niere, M.; Yakimov, A.; Redpath, P.; Khodorkovskiy, M.; Migaud, M.E.; Ziegler, M.; et al. Generation, release, and uptake of the NAD precursor nicotinic acid riboside by human cells. J. Biol. Chem. 2015, 290, 27124–27137. [Google Scholar] [CrossRef] [Green Version]
  240. Grant, R.; Berg, J.; Mestayer, R.; Braidy, N.; Bennett, J.; Broom, S.; Watson, J. A pilot study investigating changes in the human plasma and urine NAD+ metabolome during a 6 hour intravenous infusion of NAD. Front. Aging Neurosci. 2019, 11, 257. [Google Scholar] [CrossRef] [Green Version]
  241. Zarei, A.; Khazdooz, L.; Madarshahian, S.; Enayati, M.; Mosleh, I.; Lin, T.; Yan, B.; Ufheil, G.; Wooster, T.J.; Abbaspourrad, A. Synthesis, stability, and bioavailability of nicotinamide riboside trioleate chloride. Nutrients 2021, 14, 113. [Google Scholar] [CrossRef]
  242. Yang, Y.; Mohammed, F.S.; Zhang, N.; Sauve, A.A. Dihydronicotinamide riboside is a potent NAD+ concentration enhancer in vitro and in vivo. J. Biol. Chem. 2019, 294, 9295–9307. [Google Scholar] [CrossRef] [PubMed]
  243. Giroud-Gerbetant, J.; Joffraud, M.; Giner, M.P.; Cercillieux, A.; Bartova, S.; Makarov, M.V.; Zapata-Pérez, R.; Sánchez-García, J.L.; Houtkooper, R.H.; Migaud, M.E.; et al. A reduced form of nicotinamide riboside defines a new path for NAD+ biosynthesis and acts as an orally bioavailable NAD+ precursor. Mol. Metab. 2019, 30, 192–202. [Google Scholar] [CrossRef] [PubMed]
  244. Ciarlo, E.; Joffraud, M.; Hayat, F.; Giner, M.P.; Giroud-Gerbetant, J.; Sanchez-Garcia, J.L.; Rumpler, M.; Moco, S.; Migaud, M.E.; Cantó, C. Nicotinamide riboside and dihydronicotinic acid riboside synergistically increase intracellular NAD+ by generating dihydronicotinamide riboside. Nutrients 2022, 14, 2752. [Google Scholar] [CrossRef] [PubMed]
  245. Makarov, M.V.; Migaud, M.E. Syntheses and chemical properties of β-nicotinamide riboside and its analogues and derivatives. Beilstein J. Org. Chem. 2019, 15, 401–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Neamţu, A.S.; Biţă, A.; Scorei, I.R.; Rău, G.; Bejenaru, L.E.; Bejenaru, C.; Rogoveanu, O.C.; Oancea, C.N.; Radu, A.; Pisoschi, C.G.; et al. Simultaneous quantitation of nicotinamide riboside and nicotinamide in dietary supplements via HPTLC–UV with confirmation by online HPTLC–ESI–MS. Acta Chromatogr. 2020, 32, 128–133. [Google Scholar] [CrossRef]
  247. Kim, H.J.; Benner, S.A. Prebiotic stereoselective synthesis of purine and noncanonical pyrimidine nucleotide from nucleobases and phosphorylated carbohydrates. Proc. Natl. Acad. Sci. USA 2017, 114, 11315–11320. [Google Scholar] [CrossRef]
  248. Kim, H.J.; Benner, S.A. A direct prebiotic synthesis of nicotinamide nucleotide. Chemistry 2018, 24, 581–584. [Google Scholar] [CrossRef]
  249. Kim, H.J.; Benner, S.A.; Scorei, I.R. Synthesis and Stabilization of Nicotinamide Ribose and Its Derivatives. United States Patent and Trademark Office (USPTO). Provisional Patent Application No. 16160189, 15 October 2018. Available online: https://www.uspto.gov/patents (accessed on 10 May 2023).
  250. Kim, H.J.; Furukawa, Y.; Kakegawa, T.; Bita, A.; Scorei, R.; Benner, S.A. Evaporite borate-containing mineral ensembles make phosphate available and regiospecifically phosphorylate ribonucleosides: Borate as a multifaceted problem solver in prebiotic chemistry. Angew. Chem. Int. Ed Engl. 2016, 55, 15816–15820. [Google Scholar] [CrossRef]
  251. Scorei, R. Is boron a prebiotic element? A mini-review of the essentiality of boron for the appearance of life on Earth. Orig. Life Evol. Biosph. 2012, 42, 3–17. [Google Scholar] [CrossRef]
  252. Nielsen, F.H. Boron in aging and longevity. In Trace Elements and Minerals in Health and Longevity; Malavolta, M., Mocchegiani, E., Eds.; Springer: Cham, Switzerland, 2018; Volume 8, pp. 163–177. [Google Scholar] [CrossRef]
  253. Hunter, J.M.; Nemzer, B.V.; Rangavajla, N.; Biţă, A.; Rogoveanu, O.C.; Neamţu, J.; Scorei, I.R.; Bejenaru, L.E.; Rău, G.; Bejenaru, C.; et al. The fructoborates: Part of a family of naturally occurring sugar–borate complexes—Biochemistry, physiology, and impact on human health: A review. Biol. Trace Elem. Res. 2019, 188, 11–25. [Google Scholar] [CrossRef] [Green Version]
  254. Biţă, A.; Scorei, I.R.; Bălşeanu, T.A.; Ciocîlteu, M.V.; Bejenaru, C.; Radu, A.; Bejenaru, L.E.; Rău, G.; Mogoşanu, G.D.; Neamţu, J.; et al. New insights into boron essentiality in humans and animals. Int. J. Mol. Sci. 2022, 23, 9147. [Google Scholar] [CrossRef] [PubMed]
  255. Biţă, A.; Scorei, I.R.; Rangavajla, N.; Bejenaru, L.E.; Rău, G.; Bejenaru, C.; Ciocîlteu, M.V.; Dincă, L.; Neamţu, J.; Bunaciu, A.; et al. Diester chlorogenoborate complex: A new naturally occurring boron-containing compound. Inorganics 2023, 11, 112. [Google Scholar] [CrossRef]
  256. Mitruţ, I.; Scorei, I.R.; Manolea, H.O.; Biţă, A.; Mogoantă, L.; Neamţu, J.; Bejenaru, L.E.; Ciocîlteu, M.V.; Bejenaru, C.; Rău, G.; et al. Boron-containing compounds in Dentistry: A narrative review. Rom. J. Morphol. Embryol. 2022, 63, 477–483. [Google Scholar] [CrossRef] [PubMed]
  257. Neamţu, A.S.; Biţă, A.; Scorei, I.R.; Mogoşanu, G.D.; Pisoschi, C.G. Solubility, stability and degradation kinetics of nicotinamide riboside borate, a vitamin B3 derivative, in physiological solutions. Fiziologia-Physiology 2021, 31, 33–42. [Google Scholar]
Figure 1. NR borate mass spectrum: 11B (m/z 281); 10B (m/z 280). B: boron; NR: nicotinamide riboside.
Figure 1. NR borate mass spectrum: 11B (m/z 281); 10B (m/z 280). B: boron; NR: nicotinamide riboside.
Molecules 28 06078 g001
Figure 2. pH stability for NR borate.
Figure 2. pH stability for NR borate.
Molecules 28 06078 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Biţă, A.; Scorei, I.R.; Ciocîlteu, M.V.; Nicolaescu, O.E.; Pîrvu, A.S.; Bejenaru, L.E.; Rău, G.; Bejenaru, C.; Radu, A.; Neamţu, J.; et al. Nicotinamide Riboside, a Promising Vitamin B3 Derivative for Healthy Aging and Longevity: Current Research and Perspectives. Molecules 2023, 28, 6078. https://doi.org/10.3390/molecules28166078

AMA Style

Biţă A, Scorei IR, Ciocîlteu MV, Nicolaescu OE, Pîrvu AS, Bejenaru LE, Rău G, Bejenaru C, Radu A, Neamţu J, et al. Nicotinamide Riboside, a Promising Vitamin B3 Derivative for Healthy Aging and Longevity: Current Research and Perspectives. Molecules. 2023; 28(16):6078. https://doi.org/10.3390/molecules28166078

Chicago/Turabian Style

Biţă, Andrei, Ion Romulus Scorei, Maria Viorica Ciocîlteu, Oana Elena Nicolaescu, Andreea Silvia Pîrvu, Ludovic Everard Bejenaru, Gabriela Rău, Cornelia Bejenaru, Antonia Radu, Johny Neamţu, and et al. 2023. "Nicotinamide Riboside, a Promising Vitamin B3 Derivative for Healthy Aging and Longevity: Current Research and Perspectives" Molecules 28, no. 16: 6078. https://doi.org/10.3390/molecules28166078

Article Metrics

Back to TopTop