Nicotinamide adenine dinucleotide



Nicotinamide adenine dinucleotide
Other names Diphosphopyridine nucleotide (DPN+), Coenzyme I
Identifiers
CAS number 53-84-9
PubChem 925
KEGG C00003
ChEBI 13389
SMILES C1=CC(=C[N+](=C1)C2 C(C(C(O2)COP(=O)([O-])OP(=O) (O)OCC3C(C(C(O3)N4C=NC5=C 4N=CN=C5N)O)O)O)O)C(=O)N
Properties
Molecular formula P2
Molar mass 663.425
Appearance White powder
Melting point

160 °C

Hazards
Main hazards Not hazardous
NFPA 704
1
1
0
 
RTECS number UU3450000
Except where noted otherwise, data are given for
materials in their standard state
(at 25 °C, 100 kPa)

Infobox disclaimer and references

Nicotinamide adenine dinucleotide, abbreviated NAD+, is a adenosine ring, and the other containing nicotinamide.

In drug discovery.

In organisms, NAD+ can be synthesized from scratch (de novo) from the amino acids nicotinamide adenine dinucleotide phosphate (NADP+); this related coenzyme has similar chemistry to NAD+, but has different roles in metabolism.

Physical and chemical properties

Further information: Redox

Nicotinamide adenine dinucleotide is a dinucleotide since it consists of two phosphate groups through the 5' carbons.[1]

  In metabolism the compound accepts or donates electrons in redox reactions.[2] Such reactions (summarized in formula below) involve the removal of two hydrogen atoms from the reactant (R), in the form of a proton (H+). The proton is released into solution, while the reductant RH2 is oxidized and NAD+ reduced to NADH by transfer of the hydride.

RH2 + NAD+ → NADH + H+ + R

From the hydride electron pair, one electron is transferred to the positively-charged nitrogen of the nicotinamide ring of NAD+, and the second hydrogen atom transferred to the carbon atom opposite this nitrogen. The volts, which makes NADH a strong reducing agent.[3] The reaction is easily reversible, when NADH reduces another molecule and is re-oxidized to NAD+. This means the coenzyme can continuously cycle between the NAD+ and NADH forms without being consumed.[1]

In appearance, all forms of this coenzyme are white enzyme inhibitors.[5]   Both NAD+ and NADH absorb strongly in the spectrophotometer.[6]

NAD+ and NADH also differ in their fluorescence microscopy.[9]

Concentration and state in cells

In rat liver, the total amount of NAD+ and NADH is approximately 1 μmole per yeast.[2] However, over 80% is bound to proteins, so the concentration in solution is much lower.[13]

Data for other compartments in the cell are limited, although, in the diffuse across membranes.[14]

The balance between the oxidized and reduced forms of nicotinamide adenine dinucleotide is called the NAD+/NADH ratio. This ratio is an important component of what is called the redox state of a cell, a measurement that reflects both the metabolic activities and the health of cells.[15] The effects of the NAD+/NADH ratio are complex, controling the activity of several key enzymes, including NADP+/NADPH ratio is normally about 0.005, around 200 times lower than the NAD+/NADH ratio, so NADPH is the dominant form of this coenzyme.[17] These different ratios are key to the different metabolic roles of NADH and NADPH.

Biosynthesis

NAD+ is synthesized through two metabolic pathways. It is produced either in a de novo pathway from amino acids, or in salvage pathways by recycling preformed components such as nicotinamide back to NAD+.

De novo production

  Most organisms synthesize NAD+ from simple components.[2] The specific set of reactions differs among organisms, but a common feature is the generation of quinolinic acid (QA) from an amino acid - either amidated to a nicotinamide (Nam) group, forming nicotinamide adenine dinucleotide.[2]

In a further step, some NAD+ is converted into NADP+ by NAD+ kinase, which phosphorylates NAD+.[19] In most organisms, this enzyme uses ATP as the source of the phosphate group, although in bacteria such as Mycobacterium tuberculosis and in archaea such as Pyrococcus horikoshii, inorganic polyphosphate is an alternative phosphate donor.[20][21]

 

Salvage pathways

Besides assembling NAD+ de novo from simple amino acid precursors, cells also salvage preformed compounds containing nicotinamide. Although other precursors are known, the three natural compounds containing the nicotinamide ring and used in these salvage metabolic pathways are nicotinic acid (Na), nicotinamide (Nam) and nicotinamide riboside (NR).[22] The precursors are fed into the NAD(P)+ biosynthetic pathway, shown above, through adenylation and phosphoribosylation reactions.[2] These compounds are can be taken up from the diet, where the mixture of nicotinic acid and nicotinamide are called vitamin B3 or niacin. However, these compounds are also produced within cells, when the nicotinamide group is released from NAD+ in ADP-ribose transfer reactions. Indeed, the enzymes involved in these salvage pathways appear to be concentrated in the cell nucleus, which may compensate for the high level of reactions that consume NAD+ in this organelle.[23]

Despite the presence of the de novo pathway, the salvage reactions are essential in humans; a lack of niacin in the diet causes the vitamin deficiency disease pellagra.[24] This high requirement for NAD+ results from the constant consumption of the coenzyme in reactions such as posttranslational modifications, since the cycling of NAD+ between oxidized and reduced forms in redox reactions does not change the overall levels of the coenzyme.[2]

The salvage pathways used in pathogen Chlamydia trachomatis, which lacks recognizable candidates for any genes involved in the salvage or biosynthesis of both NAD+ and NADP+, and may instead salvage these coenzymes from its host.[28]

Functions

Nicotinamide adenine dinucleotide has several essential roles in sirtuins that use NAD+ to remove acetyl groups from proteins.

Oxidoreductases

 

Further information: Oxidoreductases

The main role of NAD+ in metabolism is the transfer of electrons from one redox reaction to another. This type of reaction are catalyzed by a large group of enzymes called coenzyme Q.[30] However, these enzymes are also referred to as dehydrogenases or reductases, with NADH-ubiquinone oxidoreductase commonly being called NADH dehydrogenase or sometimes coenzyme Q reductase.[31]

When bound to a protein, NAD+ and NADH are usually held within a amino acid metabolism have recently been discovered that bind the coenzyme, but lack this motif.[34]

Despite this similarity in how proteins bind coenzymes, enzymes almost always show a high level of specificity for either NAD+ or NADP+.[35] This specificity reflects the distinct metabolic roles of the two coenzymes, and is the result of distinct sets of ionic bond is formed between a basic amino acid side chain and the acidic phosphate group of NADP+. Conversely, in NAD-dependent enzymes the charge in this pocket is reversed, preventing NADP+ from binding. However, there are a few exceptions to this general rule, and enzymes such aldose reductase, glucose-6-phosphate dehydrogenase, and methylenetetrahydrofolate reductase can use both coenzymes in some species.[36]

Role in redox metabolism

 

Further information: Oxidative phosphorylation

The redox reactions catalyzed by oxidoreductases are vital in all parts of metabolism, but one particularly important area where these reactions occur is in the release of energy from nutrients. Here, reduced compounds such as chloroplasts.[38]

Since both the oxidized and reduced forms of nicotinamide adenine dinucleotide are used in these linked sets of reactions, the cell maintains approximately equal concentrations of NAD+ and NADH; the high NAD+/NADH ratio allows this coenzyme to act as both an oxidizing and a reducing agent.[39] In contrast, the main function of NADP+ is as a reducing agent in photosynthesis. Since NADPH is needed to drive redox reactions as a strong reducing agent, the NADP+/NADPH ratio is kept very low.[39]

Although it is important in catabolism, NADH is also used in anabolic reactions, such as proton-motive force to run part of the electron transport chain in reverse, generating NADH.[42]

Non-redox roles

The coenzyme NAD+ is also consumed in ADP-ribose transfer reactions. For example, enzymes called telomere maintenance.[47] In addition to these functions within the cell, a group of extracellular ADP-ribosyltransferases has recently been discovered, but their functions remain obscure.[48]   Another function of this coenzyme in cell signaling is as a precursor of ryanodine receptors, which are located in the membranes of organelles, such as the endoplasmic reticulum.[51]

NAD+ is also consumed by transcription through deacetylating histones and altering nucleosome structure.[53] These activities of sirtuins are particularly interesting due to their importance in the regulation of aging.[54]

Other NAD-dependent enzymes include bacterial ATP to form the DNA-AMP intermediate.[56]

Pharmacology

The enzymes that make and use NAD+ and NADH are important in both current enzyme inhibitors or activators based on its structure that change the activity of NAD-dependent enzymes, and by trying to inhibit NAD+ biosynthesis.[57]

The coenzyme NAD+ is not itself currently used as a treatment for any disease. However, it is potentially useful in the therapy of neurodegenerative diseases such as Alzheimer's and Parkinson disease.[2] Evidence for these applications is mixed; studies in mice are promising,[58] whereas a placebo-controlled clinical trial failed to show any effect.[59] NAD+ is also a direct target of the drug free radical form.[60] This radical then reacts with NADH, to produce adducts that are very potent inhibitors of the enzymes enoyl-acyl carrier protein reductase,[61] and dihydrofolate reductase.[62]

Since a large number of oxidoreductases use NAD+ and NADH as substrates, and bind them using a highly-conserved structural motif, the idea that inhibitors based on NAD+ could be specific to one enzyme is surprizing.[63] However, this can be possible: for example, inhibitors based on the compounds resveratrol increase the activity of these enzymes, which may be important in their ability to delay aging in both vertebrate,[65] and invertebrate model organisms.[66][67]

Due to the differences in the antibiotics.[68][69] For example, the enzyme nicotinamidase, which converts nicotinamide to nicotinic acid, is a target for drug design, as this enzyme is absent in humans but present in yeast and bacteria.[25]

History

Further information: History of biochemistry

  The coenzyme NAD+ was first discovered by the British biochemists Arthur Harden and William Youndin in 1906.[70] They noticed that adding boiled and filtered nucleotide sugar phosphate by Hans von Euler-Chelpin.[71] In 1936, the German scientist Otto Heinrich Warburg showed the function of the nucleotide coenzyme in hydride transfer and identified the nicotinamide portion as the site of redox reactions.[72]

The source of nicotinamide was identified in 1938, when Arthur Kornberg made another important contribution towards understanding NAD+ metabolism, by being the first to detect an enzyme in the biosynthetic pathway.[75] Subsequently, in 1949, the American biochemists Morris Friedkin and Albert L. Lehninger proved that NADH linked metabolic pathways such as the citric acid cycle with the synthesis of ATP in oxidative phosphorylation.[76] Finally, in 1959, Jack Preiss and Philip Handler discovered the intermediates and enzymes involved in the biosynthesis of NAD+;[77][78] consequently, de novo synthesis is often called the Preiss-Handler pathway in their honor.

The non-redox roles of NAD(P) are a recent discovery.[1] The first of these functions to be identified was the use of NAD+ as the ADP-ribose donor in ADP-ribosylation reactions, observed in the early 1960s.[79] Later studies in the 1980s and 1990s revealed the activities of NAD+ and NADP+ metabolites in cell signaling - such as the action of sirtuins in 2000, by Shin-ichiro Imai and coworkers at the Massachusetts Institute of Technology.[81]

See also

References

  1. ^ a b c Pollak, N; Dölle C, Ziegler M (2007). "The power to reduce: pyridine nucleotides—small molecules with a multitude of functions". Biochem. J. 402 (2): 205–18. PMID 17295611. Retrieved on 2007-12-23.
  2. ^ a b c d e f g Belenky P; Bogan KL, Brenner C (2007). "NAD+ metabolism in health and disease". Trends Biochem. Sci. 32 (1): 12–9. PMID 17161604. Retrieved on 2007-12-23.
  3. ^ Unden G; Bongaerts J (1997). "Alternative respiratory pathways of Escherichia coli: energetics and transcriptional regulation in response to electron acceptors". Biochim. Biophys. Acta 1320 (3): 217–34. PMID 9230919.
  4. ^ The Merck Index tenth edition, (Merck & Co. Ltd 1983) ISBN 9-11-91027-1 p 909
  5. ^ Biellmann JF, Lapinte C, Haid E, Weimann G (1979). "Structure of lactate dehydrogenase inhibitor generated from coenzyme". Biochemistry 18 (7): 1212–7. PMID 218616.
  6. ^ a b Dawson MC (ed) Data for biochemical research third edition, (Oxford scientific publications, 1987) ISBN 0-19-855358-7 p 122
  7. ^ a b Lakowicz JR, Szmacinski H, Nowaczyk K, Johnson ML (1992). "Fluorescence lifetime imaging of free and protein-bound NADH". Proc. Natl. Acad. Sci. U.S.A. 89 (4): 1271–5. PMID 1741380.
  8. ^ Jameson DM, Thomas V, Zhou DM (1989). "Time-resolved fluorescence studies on NADH bound to mitochondrial malate dehydrogenase". Biochim. Biophys. Acta 994 (2): 187–90. PMID 2910350.
  9. ^ Kasimova MR, Grigiene J, Krab K, et al. (2006). "The free NADH concentration is kept constant in plant mitochondria under different metabolic conditions". Plant Cell 18 (3): 688–98. PMID 16461578.
  10. ^ Reiss PD, Zuurendonk PF, Veech RL (1984). "Measurement of tissue purine, pyrimidine, and other nucleotides by radial compression high-performance liquid chromatography". Anal. Biochem. 140 (1): 162–71. PMID 6486402.
  11. ^ Yamada K, Hara N, Shibata T, Osago H, Tsuchiya M (2006). "The simultaneous measurement of nicotinamide adenine dinucleotide and related compounds by liquid chromatography/electrospray ionization tandem mass spectrometry". Anal. Biochem. 352 (2): 282–5. PMID 16574057.
  12. ^ a b Yang H, Yang T, Baur JA, Perez E, Matsui T, Carmona JJ, Lamming DW, Souza-Pinto NC, Bohr VA, Rosenzweig A, de Cabo R, Sauve AA, Sinclair DA. (2007). "Nutrient-Sensitive Mitochondrial NAD+ Levels Dictate Cell Survival". Cell 130: 1095–107.
  13. ^ Blinova K, Carroll S, Bose S, et al (2005). "Distribution of mitochondrial NADH fluorescence lifetimes: steady-state kinetics of matrix NADH interactions". Biochemistry 44 (7): 2585–94. PMID 15709771.
  14. ^ Todisco S, Agrimi G, Castegna A, Palmieri F (2006). "Identification of the mitochondrial NAD+ transporter in Saccharomyces cerevisiae". J. Biol. Chem. 281 (3): 1524–31. PMID 16291748.
  15. ^ Schafer F, Buettner G (2001). "Redox environment of the cell as viewed through the redox state of the glutathione disulfide/glutathione couple". Free Radic Biol Med 30 (11): 1191–212. PMID 11368918.
  16. ^ a b c Lin SJ, Guarente L (2003). "Nicotinamide adenine dinucleotide, a metabolic regulator of transcription, longevity and disease". Curr. Opin. Cell Biol. 15 (2): 241–6. PMID 12648681.
  17. ^ Veech RL, Eggleston LV, Krebs HA (1969). "The redox state of free nicotinamide-adenine dinucleotide phosphate in the cytoplasm of rat liver". Biochem. J. 115 (4): 609–19. PMID 4391039.
  18. ^ Katoh A, Uenohara K, Akita M, Hashimoto T (2006). "Early steps in the biosynthesis of NAD in Arabidopsis start with aspartate and occur in the plastid". Plant Physiol. 141 (3): 851–7. PMID 16698895.
  19. ^ Magni G, Orsomando G, Raffaelli N (2006). "Structural and functional properties of NAD kinase, a key enzyme in NADP biosynthesis". Mini reviews in medicinal chemistry 6 (7): 739–46. PMID 16842123.
  20. ^ Sakuraba H, Kawakami R, Ohshima T (2005). "First archaeal inorganic polyphosphate/ATP-dependent NAD kinase, from hyperthermophilic archaeon Pyrococcus horikoshii: cloning, expression, and characterization". Appl. Environ. Microbiol. 71 (8): 4352–8. PMID 16085824.
  21. ^ Raffaelli N, Finaurini L, Mazzola F, et al (2004). "Characterization of Mycobacterium tuberculosis NAD kinase: functional analysis of the full-length enzyme by site-directed mutagenesis". Biochemistry 43 (23): 7610–7. PMID 15182203.
  22. ^ Tempel W, Rabeh WM, Bogan KL, et al (2007). "Nicotinamide riboside kinase structures reveal new pathways to NAD+". PLoS Biol. 5 (10): e263. PMID 17914902.
  23. ^ Anderson RM, Bitterman KJ, Wood JG, et al (2002). "Manipulation of a nuclear NAD+ salvage pathway delays aging without altering steady-state NAD+ levels". J. Biol. Chem. 277 (21): 18881–90. PMID 11884393.
  24. ^ Henderson LM (1983). "Niacin". Annu. Rev. Nutr. 3: 289–307. PMID 6357238.
  25. ^ a b Rongvaux A, Andris F, Van Gool F, Leo O (2003). "Reconstructing eukaryotic NAD metabolism". Bioessays 25 (7): 683–90. PMID 12815723.
  26. ^ Ma B, Pan SJ, Zupancic ML, Cormack BP (2007). "Assimilation of NAD(+) precursors in Candida glabrata". Mol. Microbiol. 66 (1): 14–25. doi:doi:10.1111/j.1365-2958.2007.05886.x.
  27. ^ Reidl J, Schlör S, Kraiss A, Schmidt-Brauns J, Kemmer G, Soleva E (2000). "NADP and NAD utilization in Haemophilus influenzae". Mol. Microbiol. 35 (6): 1573–81. PMID 10760156.
  28. ^ Gerdes SY, Scholle MD, D'Souza M, et al (2002). "From genetic footprinting to antimicrobial drug targets: examples in cofactor biosynthetic pathways". J. Bacteriol. 184 (16): 4555–72. PMID 12142426.
  29. ^ Senkovich O, Speed H, Grigorian A, et al (2005). "Crystallization of three key glycolytic enzymes of the opportunistic pathogen Cryptosporidium parvum". Biochim. Biophys. Acta 1750 (2): 166–72. PMID 15953771.
  30. ^ Enzyme Nomenclature, Recommendations for enzyme names from the Nomenclature Committee of the International Union of Biochemistry and Molecular Biology. Retrieved on 2007-12-06.
  31. ^ NiceZyme View of ENZYME: EC 1.6.5.3. Expasy. Retrieved on 2007-12-16.
  32. ^ Lesk AM (1995). "NAD-binding domains of dehydrogenases". Curr. Opin. Struct. Biol. 5 (6): 775–83. PMID 8749365.
  33. ^ a b Rao S, Rossmann M (1973). "Comparison of super-secondary structures in proteins". J Mol Biol 76 (2): 241–56. PMID 4737475.
  34. ^ Goto M, Muramatsu H, Mihara H, et al (2005). "Crystal structures of Delta1-piperideine-2-carboxylate/Delta1-pyrroline-2-carboxylate reductase belonging to a new family of NAD(P)H-dependent oxidoreductases: conformational change, substrate recognition, and stereochemistry of the reaction". J. Biol. Chem. 280 (49): 40875–84. PMID 16192274.
  35. ^ Carugo O, Argos P (1997). "NADP-dependent enzymes. I: Conserved stereochemistry of cofactor binding". Proteins 28 (1): 10–28. PMID 9144787.
  36. ^ Vickers TJ, Orsomando G, de la Garza RD, et al (2006). "Biochemical and genetic analysis of methylenetetrahydrofolate reductase in Leishmania metabolism and virulence". J. Biol. Chem. 281 (50): 38150–8. PMID 17032644.
  37. ^ Rich PR (2003). "The molecular machinery of Keilin's respiratory chain". Biochem. Soc. Trans. 31 (Pt 6): 1095–105. PMID 14641005.
  38. ^ Heineke D, Riens B, Grosse H, et al (1991). "Redox Transfer across the Inner Chloroplast Envelope Membrane". Plant Physiol 95 (4): 1131–1137. PMID 16668101.
  39. ^ a b Nicholls DG; Ferguson SJ (2002). Bioenergetics 3, 1st ed, Academic Press. ISBN 0-125-18121-3. 
  40. ^ Sistare FD, Haynes RC (1985). "The interaction between the cytosolic pyridine nucleotide redox potential and gluconeogenesis from lactate/pyruvate in isolated rat hepatocytes. Implications for investigations of hormone action". J. Biol. Chem. 260 (23): 12748–53. PMID 4044607.
  41. ^ Freitag A, Bock E (1990). "Energy conservation in Nitrobacter". FEMS Microbiology Letters 66 (1–3): 157&ndash:62. doi:10.1111/j.1574-6968.1990.tb03989.x.
  42. ^ Starkenburg SR, Chain PS, Sayavedra-Soto LA, et al (2006). "Genome sequence of the chemolithoautotrophic nitrite-oxidizing bacterium Nitrobacter winogradskyi Nb-255". Appl. Environ. Microbiol. 72 (3): 2050–63. PMID 16517654.
  43. ^ Ziegler M (2000). "New functions of a long-known molecule. Emerging roles of NAD in cellular signaling". Eur. J. Biochem. 267 (6): 1550–64. PMID 10712584.
  44. ^ a b Diefenbach J, Bürkle A (2005). "Introduction to poly(ADP-ribose) metabolism". Cell. Mol. Life Sci. 62 (7-8): 721–30. PMID 15868397.
  45. ^ Berger F, Ramírez-Hernández MH, Ziegler M (2004). "The new life of a centenarian: signaling functions of NAD(P)". Trends Biochem. Sci. 29 (3): 111–8. PMID 15003268.
  46. ^ Corda D, Di Girolamo M (2003). "Functional aspects of protein mono-ADP-ribosylation". EMBO J. 22 (9): 1953–8. PMID 12727863.
  47. ^ a b Burkle A (2005). "Poly(ADP-ribose). The most elaborate metabolite of NAD+". FEBS J. 272 (18): 4576–89. PMID 16156780.
  48. ^ Seman M, Adriouch S, Haag F, Koch-Nolte F (2004). "Ecto-ADP-ribosyltransferases (ARTs): emerging actors in cell communication and signaling". Curr. Med. Chem. 11 (7): 857–72. PMID 15078170.
  49. ^ Guse AH (2004). "Biochemistry, biology, and pharmacology of cyclic adenosine diphosphoribose (cADPR)". Curr. Med. Chem. 11 (7): 847–55. PMID 15078169.
  50. ^ Guse AH (2004). "Regulation of calcium signaling by the second messenger cyclic adenosine diphosphoribose (cADPR)". Curr. Mol. Med. 4 (3): 239–48. PMID 15101682.
  51. ^ Guse AH (2005). "Second messenger function and the structure-activity relationship of cyclic adenosine diphosphoribose (cADPR)". FEBS J. 272 (18): 4590–7. PMID 16156781.
  52. ^ North B, Verdin E (2004). "Sirtuins: Sir2-related NAD-dependent protein deacetylases". Genome Biol 5 (5): 224. PMID 15128440.
  53. ^ Blander G, Guarente L (2004). "The Sir2 family of protein deacetylases". Annu. Rev. Biochem. 73: 417–35. PMID 15189148.
  54. ^ Trapp J, Jung M (2006). "The role of NAD+ dependent histone deacetylases (sirtuins) in ageing". Curr Drug Targets 7 (11): 1553–60. PMID 17100594.
  55. ^ Wilkinson A, Day J, Bowater R (2001). "Bacterial DNA ligases". Mol. Microbiol. 40 (6): 1241–8. PMID 11442824.
  56. ^ Schär P, Herrmann G, Daly G, Lindahl T (1997). "A newly identified DNA ligase of Saccharomyces cerevisiae involved in RAD52-independent repair of DNA double-strand breaks". Genes and Development 11 (15): 1912–24. PMID 9271115.
  57. ^ Khan JA, Forouhar F, Tao X, Tong L (2007). "Nicotinamide adenine dinucleotide metabolism as an attractive target for drug discovery". Expert Opin. Ther. Targets 11 (5): 695–705. PMID 17465726.
  58. ^ Kaneko S, Wang J, Kaneko M, et al (2006). "Protecting axonal degeneration by increasing nicotinamide adenine dinucleotide levels in experimental autoimmune encephalomyelitis models". J. Neurosci. 26 (38): 9794–804. PMID 16988050.
  59. ^ Swerdlow RH (1998). "Is NADH effective in the treatment of Parkinson's disease?". Drugs Aging 13 (4): 263–8. PMID 9805207.
  60. ^ Timmins GS, Deretic V (2006). "Mechanisms of action of isoniazid". Mol. Microbiol. 62 (5): 1220–7. PMID 17074073.
  61. ^ Rawat R, Whitty A, Tonge PJ (2003). "The isoniazid-NAD adduct is a slow, tight-binding inhibitor of InhA, the Mycobacterium tuberculosis enoyl reductase: adduct affinity and drug resistance". Proc. Natl. Acad. Sci. U.S.A. 100 (24): 13881–6. PMID 14623976.
  62. ^ Argyrou A, Vetting MW, Aladegbami B, Blanchard JS (2006). "Mycobacterium tuberculosis dihydrofolate reductase is a target for isoniazid". Nat. Struct. Mol. Biol. 13 (5): 408–13. PMID 16648861.
  63. ^ a b Pankiewicz KW, Patterson SE, Black PL, et al (2004). "Cofactor mimics as selective inhibitors of NAD-dependent inosine monophosphate dehydrogenase (IMPDH)--the major therapeutic target". Curr. Med. Chem. 11 (7): 887–900. PMID 15083807.
  64. ^ Franchetti P, Grifantini M (1999). "Nucleoside and non-nucleoside IMP dehydrogenase inhibitors as antitumor and antiviral agents". Curr. Med. Chem. 6 (7): 599–614. PMID 10390603.
  65. ^ Valenzano DR, Terzibasi E, Genade T, Cattaneo A, Domenici L, Cellerino A (2006). "Resveratrol prolongs lifespan and retards the onset of age-related markers in a short-lived vertebrate". Curr. Biol. 16 (3): 296–300. PMID 16461283.
  66. ^ Howitz KT, Bitterman KJ, Cohen HY, et al (2003). "Small molecule activators of sirtuins extend Saccharomyces cerevisiae lifespan". Nature 425 (6954): 191–6. PMID 12939617.
  67. ^ Wood JG, Rogina B, Lavu S, et al (2004). "Sirtuin activators mimic caloric restriction and delay ageing in metazoans". Nature 430 (7000): 686–9. PMID 15254550.
  68. ^ Rizzi M, Schindelin H (2002). "Structural biology of enzymes involved in NAD and molybdenum cofactor biosynthesis". Curr. Opin. Struct. Biol. 12 (6): 709–20. PMID 12504674.
  69. ^ Begley TP, Kinsland C, Mehl RA, Osterman A, Dorrestein P (2001). "The biosynthesis of nicotinamide adenine dinucleotides in bacteria". Vitam. Horm. 61: 103–19. PMID 11153263.
  70. ^ Harden, A; Young, WJ (October 1906). "The Alcoholic Ferment of Yeast-Juice" 78: pp. 369–375.
  71. ^ Fermentation of sugars and fermentative enzymes. Nobel Lecture, 23 May 1930. Nobel Foundation. Retrieved on 2007-09-30.
  72. ^ Warburg O, Christian W. (1936). "Pyridin, the hydrogen-transferring component of the fermentation enzymes (pyridine nucleotide)". Biochemische Zeitschrift 287.
  73. ^ Elvehjem CA, Madden RJ, Strong FM, Woolley DW. (1938). "The isolation and identification of the anti-black tongue factor". J. Biol. Chem. 123 (1): 137–49.
  74. ^ Axelrod AE, Madden RJ, Elvehjem CA, (1939). "The effect of a nicotinic acid deficiency upon the coenzyme I content of animal tissues". J. Biol. Chem. 131 (1): 85–93.
  75. ^ Kornberg, A. (1948). "The participation of inorganic pyrophosphate in the reversible enzymatic synthesis of diphosphopyridine nucleotide.". J. Biol. Chem. 176 (3): 1475–76.
  76. ^ Friedkin M, Lehninger AL. (1949). "Esterification of inorganic phosphate coupled to electron transport between dihydrodiphosphopyridine nucleotide and oxygen". J. Biol. Chem. 178 (2): 611–23.
  77. ^ Preiss J, Handler P. (1958). "Biosynthesis of diphosphopyridine nucleotide. I. Identification of intermediates". J. Biol. Chem. 233 (2): 488–92. PMID 13563526.
  78. ^ Preiss J, Handler P. (1958). "Biosynthesis of diphosphopyridine nucleotide. II. Enzymatic aspects". J. Biol. Chem. 233 (2): 493–500. PMID 13563527.
  79. ^ Chambon P, Weill JD, Mandel P (1963). "Nicotinamide mononucleotide activation of new DNA-dependent polyadenylic acid synthesizing nuclear enzyme". Biochem. Biophys. Res. Commun. 11: 39–43. PMID 14019961.
  80. ^ Clapper DL, Walseth TF, Dargie PJ, Lee HC (1987). "Pyridine nucleotide metabolites stimulate calcium release from sea urchin egg microsomes desensitized to inositol trisphosphate". J. Biol. Chem. 262 (20): 9561–8. PMID 3496336.
  81. ^ Imai S, Armstrong CM, Kaeberlein M, Guarente L (2000). "Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase". Nature 403 (6771): 795–800. PMID 10693811.

Further reading

Function

  • Nelson DL; Cox MM (2004). Lehninger Principles of Biochemistry, 4th ed, W. H. Freeman. ISBN 0-716-74339-6. 
  • Bugg, T (2004). Introduction to Enzyme and Coenzyme Chemistry, 2nd ed, Blackwell Publishing Limited. ISBN 1-40511-452-5. 
  • Lee HC (2002). Cyclic ADP-Ribose and NAADP: Structure, Metabolism and Functions. Kluwer Academic Publishers. ISBN 1-40207-281-3. 

History

  • New Beer in an Old Bottle: Eduard Buchner and the Growth of Biochemical Knowledge, edited by Athel Cornish-Bowden and published by Universitat de València (1997): ISBN 84-370-3328-4, A history of early enzymology.
  • Williams, Henry Smith, 1863–1943. A History of Science: in Five Volumes. Volume IV: Modern Development of the Chemical and Biological Sciences, A textbook from the 19th century.


 
This article is licensed under the GNU Free Documentation License. It uses material from the Wikipedia article "Nicotinamide_adenine_dinucleotide". A list of authors is available in Wikipedia.