Hostname: page-component-586b7cd67f-r5fsc Total loading time: 0 Render date: 2024-11-27T13:48:29.331Z Has data issue: false hasContentIssue false

The role of dietary niacin intake and the adenosine-5′-diphosphate-ribosyl cyclase enzyme CD38 in spatial learning ability: is cyclic adenosine diphosphate ribose the link between diet and behaviour?

Published online by Cambridge University Press:  01 June 2008

Genevieve S. Young*
Affiliation:
Department of Human Health and Nutritional Sciences, University of Guelph, Guelph, Ontario, Canada
James B. Kirkland
Affiliation:
Department of Human Health and Nutritional Sciences, University of Guelph, Guelph, Ontario, Canada
*
*Corresponding author: Dr Genevieve Young, fax +1 519 763 5902, email [email protected]
Rights & Permissions [Opens in a new window]

Abstract

The pyridine nucleotide NAD+ is derived from dietary niacin and serves as the substrate for the synthesis of cyclic ADP-ribose (cADPR), an intracellular Ca signalling molecule that plays an important role in synaptic plasticity in the hippocampus, a region of the brain involved in spatial learning. cADPR is formed in part via the activity of the ADP-ribosyl cyclase enzyme CD38, which is widespread throughout the brain. In the present review, current evidence of the relationship between dietary niacin and behaviour is presented following investigations of the effect of niacin deficiency, pharmacological nicotinamide supplementation and CD38 gene deletion on brain nucleotides and spatial learning ability in mice and rats. In young male rats, both niacin deficiency and nicotinamide supplementation significantly altered brain NAD+ and cADPR, both of which were inversely correlated with spatial learning ability. These results were consistent across three different models of niacin deficiency (pair feeding, partially restricted feeding and niacin recovery). Similar changes in spatial learning ability were observed in Cd38− / −  mice, which also showed decreases in brain cADPR. These findings suggest an inverse relationship between spatial learning ability, dietary niacin intake and cADPR, although a direct link between cADPR and spatial learning ability is still missing. Dietary niacin may therefore play a role in the molecular events regulating learning performance, and further investigations of niacin intake, CD38 and cADPR may help identify potential molecular targets for clinical intervention to enhance learning and prevent or reverse cognitive decline.

Type
Research Article
Copyright
Copyright © The Authors 2008

Introduction

Niacin is the term used to describe vitamers including nicotinamide (pyridine-3-carboxyamide; Fig. 1(a)), nicotinic acid (pyridine-3-carboxylic acid; Fig. 1(b)), and a variety of pyridine nucleotide structures, such as NAD (Fig. 1(c)) and NADP (Fig. 1(d))(Reference Kirkland, Rucker, Zempleni, Suttie and McCormick1). NAD is synthesised from nicotinic acid via the Preiss–Handler pathway(Reference Preiss and Handler2) or from nicotinamide via the Dietrich pathway(Reference Dietrich, Fuller, Yero and Martinez3). Niacin, through NAD, can also be formed from the essential amino acid tryptophan. When niacin is derived from tryptophan, approximately 1 mg of the vitamin is formed from 60 mg of the amino acid, although the efficiency of conversion is affected by factors such as tryptophan and niacin intake and the amino acid, carbohydrate, vitamin B6 and fat content of the diet(Reference Shibata, Mushiage, Kondo, Hayakawa and Tsuge4, Reference Shibata5). As NAD+, niacin is involved in a number of biochemical processes, including energy metabolism (redox reactions), protein modification by mono- and poly-(ADP-ribose) polymerases, and synthesis of intracellular Ca signalling molecules(Reference Kirkland, Rucker, Zempleni, Suttie and McCormick1).

Fig. 1 Chemical structures of niacin compounds: (a) nicotinamide; (b) nicotinic acid; (c) nicotinamide adenine dinucleotide (NAD+); (d) nicotinamide adenine dinucleotide phosphate (NADP+).

There is a long history of research concerning niacin status and brain function. Niacin deficiency in humans causes pellagra, which is characterised by sun-sensitivity and dementia. Neurological changes in pellagra patients begin peripherally, with signs such as muscle weakness, twitching and burning feelings in the extremities and altered gait(Reference Buniva and Carpenter6). Early psychological changes include depression and apprehension, but these progress to more severe changes, such as vertigo, loss of memory, deep depression, paranoia and delirium, hallucinations and violent behaviour(Reference Spies, Bean, Ashe and Carpenter7), similar to schizophrenia(Reference Hoffer8). While there are pathological changes in the spinal cord in advanced pellagra, there is a striking recovery of psychological function when insane pellagra patients are treated with nicotinic acid, with a disappearance of many symptoms in 1–2 d(Reference Spies, Bean, Ashe and Carpenter7). These observations suggest that a compound derived from niacin is involved in neural signalling pathways. The recent discovery that the intracellular Ca signalling molecule cyclic ADP-ribose (cADPR) is derived from NAD+ (Fig. 2)(Reference Lee, Walseth, Bratt, Hayes and Clapper9) suggests that cADPR might be the link between niacin status and behaviour. cADPR is involved in synaptic plasticity in the hippocampus(Reference Reyes-Harde, Potter, Galione and Stanton10, Reference Reyes-Harde, Empson, Potter, Galione and Stanton11), a region of the brain that regulates spatial learning(Reference Redish and Touretzky12). As NAD+ is derived from dietary niacin, cADPR levels might be expected to change with dietary niacin intake. The relationship between niacin, cADPR and hippocampal synaptic plasticity is the basis for the investigations described in the present review.

Fig. 2 Structure and origin of cyclic adenosine diphosphate ribose.

Discussion

Intracellular calcium

Modulation of intracellular Ca ion concentration is a universal mechanism by which extracellular signals are transduced into an intracellular response(Reference Berridge, Lipp and Bootman13). Ca levels inside the cell are controlled by both ion influx through channels in the plasma membrane and by release from intracellular stores, and Ca channels in plasma and organelle membranes open in response to extracellular signals in a spatial and temporally specific pattern to cause both local and global increases in intracellular ion concentration(Reference Carafoli, Santella, Branca and Brini14). Intracellular Ca stores in the cell include (1) the endoplasmic reticulum(Reference Meldolesi and Pozzan15), (2) the mitochondria(Reference Pozzan, Magalhaes and Rizzuto16Reference Collins, Berridge, Lipp and Bootman18), (3) the nuclear envelope(Reference Malviya, Rogue and Vincendon19Reference Gerasimenko, Gerasimenko, Tepikin and Petersen21), (4) the Golgi apparatus(Reference Pinton, Pozzan and Rizzuto22), (5) secretory granules(Reference Yoo23) and (6) endosomes(Reference Gerasimenko, Tepikin, Petersen and Gerasimenko24). There are two other intracellular Ca mobilising molecules in addition to cADPR: nicotinic acid adenine dinucleotide phosphate (NAADP), which is formed from phosphorylated NAD(Reference Lee and Aarhus25), and d-myo-inositol 1,4,5-triphosphate (IP3)(Reference Streb, Irvine, Berridge and Schulz26). A summary of the characteristics of IP3, cADPR and NAADP is presented in Table 1(Reference Lee, Walseth, Bratt, Hayes and Clapper9, Reference Lee and Aarhus25, Reference Mikoshiba, Furuichi and Miyawaki27Reference Aarhus, Graeff, Dickey, Walseth and Lee44). Multiplicity of Ca signalling pathways may serve several functions, including redundancy to ensure that Ca signalling occurs and variation in the spatial and temporal Ca response(Reference da Silva and Guse45). Other compounds such as lysophosphatidic acid(Reference Melendez and Allen46), sphingosine 1-phosphate(Reference Young and Nahorski47) and ADP-ribose(Reference Lee48) are also involved in intracellular Ca mobilisation, although these do not necessarily function as second messengers.

Table 1 Characteristics of inositol 1,4,5-triphosphate (IP3), cyclic adenosine diphosphate ribose (cADPR) and nicotinic acid adenine dinucleotide phosphate (NAADP)

ER, endoplasmic reticulum; SR, sarcoplasmic reticulum; IP3R, IP3 receptor; RyR, ryanodine receptor; NAADPR, NAADP receptor; CICR, Ca-induced Ca release.

Ca signalling plays a crucial role in regulating many neuronal processes. As reviewed by Berridge(Reference Berridge49), N- and P/Q-type voltage-activated channels are localised in synaptic terminals, where they regulate neurotransmitter release by generating a local Ca transient which activates synaptotagmin and triggers exocytosis. L-type voltage-activated channels are found on the cell body and proximal dendrites and regulate gene transcription. Synaptic plasticity is thought to be mediated by Ca entry through both voltage- and receptor-operated channels, and by release from IP3 receptors and ryanodine receptors (RyR). As will be discussed later, modulation of intracellular Ca is required for both long-term potentiation (LTP) and long-term depression (LTD), cellular mechanisms which are thought to contribute to learning and memory(Reference Franks and Sejnowski50). Similar to neurons, astrocytes also regulate intracellular function through generation of Ca signals; as well, they control the function of neighbouring neurons through global Ca transients(Reference Scemes51).

Cyclic adenosine diphosphate ribose

cADPR has been found to mobilise intracellular Ca in numerous cell types, including protozoa, and those of plants, animals and man(Reference Guse52). cADPR mobilises Ca ions via Ca-induced Ca release, whereby the Ca2+-releasing mechanism is sensitised by the addition of Ca2+(Reference Galione, Lee and Busa53). cADPR synthesis is stimulated by cGMP(Reference Galione, White, Willmott, Turner, Potter and Watson54). cADPR also activates extracellular Ca influx(Reference Guse, Berg, da Silva, Potter and Mayr55, Reference Partida-Sánchez, Cockayne and Monard56). The function of cADPR has been investigated in a wide range of cell types, and a summary of the intracellular effects of cADPR is presented in Table 2 (summarised in part from Guse(Reference Guse52)). Table 2(Reference Gerasimenko, Gerasimenko, Tepikin and Petersen21, Reference Partida-Sánchez, Cockayne and Monard56Reference Yusufi, Cheng, Thompson, Dousa, Warner, Walker and Grande92) shows that cADPR administration brings about a variety of changes in neurons, including neurotransmitter release.

Table 2 Intracellular effects of cyclic adenosine diphosphate ribose

SR, sarcoplasmic reticulum; CICR, Ca-induced Ca release.

The principal target of cADPR is the RyR(Reference Galione, Lee and Busa53). cADPR might bind directly to the RyR, or an additional binding protein might be required. The binding protein FKPB 12·6 has been found to act as a cADPR-binding protein in several cell types, including pancreatic islets(Reference Noguchi, Takasawa, Nata, Tohgo, Kato, Ikehata, Yonekura and Okamoto84) and smooth muscle cells(Reference Tang, Chen, Zou, Campbell and Li66, Reference Li, Tang, Valdivia, Zou and Campbell64). Binding of cADPR to a target protein might cause release of the protein from the RyR, allowing opening of the RyR Ca channel(Reference Guse93). The Ca-binding protein calmodulin is involved in cADPR-mediated Ca release from the RyR(Reference Lee, Aarhus, Graeff, Gurnack and Walseth94), and tyrosine phosphorylation of the RyR increases cADPR-mediated Ca release(Reference Guse, Tsygankov, Weber and Mayr95). RyR, which are primarily found on the endoplasmic reticulum/sarcoplasmic reticulum (ER/SR) membrane, are not evenly distributed throughout the cell and distribution patterns vary across cell types(Reference Guse93). In addition to the ER/SR, RyR are also found in mitochondria(Reference Beutner, Sharma, Giovannucci, Yule and Sheu96) and in the nuclear envelope(Reference Khoo and Chang97, Reference Adebanjo, Anandatheerthavarada and Koval98). Of the three RyR isoforms, cADPR has been shown to bind to type II and III(Reference Galione, White, Willmott, Turner, Potter and Watson54, Reference Schwarzmann, Kunerth, Weber, Mayr and Guse99). An alternative mechanism has been proposed for cADPR whereby cADPR promotes refilling of depleted Ca stores rather than acting on RyR to induce Ca release(Reference Lukyanenko, Gyorke, Wiesner and Gyorke100). The precise target of cADPR and any associated binding proteins is still not well understood.

CD38

The synthesis of cADPR is catalysed by the ADP-ribosyl cyclase family of enzymes. These include: CD38, a type II ectoenzyme that is also expressed intracellularly(Reference Lee, Graeff and Walseth39); BST-1, also known as CD157, a bone marrow stromal cell-surface antigen(Reference Itoh, Ishihara, Tomizawa, Tanaka, Kobune, Ishikawa, Kaisho and Hirano101); a soluble cyclase characterised from the ovotestes of the Aplysia mollusk and of dogs(Reference Lee, Graeff and Walseth39); a membrane-bound cyclase from canine spleen(Reference Kim, Jacobson and Jacobson102); a membrane-bound cyclase from mouse brain(Reference Ceni, Pochon, Villaz, Muller-Steffner, Schuber, Baratier, De Waard, Ronjat and Moutin103). The ADP-ribosyl cyclase enzymes are multifunctional, catalysing three reactions: (1) cyclisation of NAD+ (ADP-ribosyl cyclase activity), (2) hydrolysis of cADPR to ADP-ribose (cADPR hydrolase activity) and (3) hydrolysis of NAD+ to ADP-ribose (NAD+ hydrolase activity)(Reference Lee, Graeff and Walseth39). ADP-ribosyl cyclase enzymes also catalyse the exchange of nicotinamide and nicotinic acid to form NAADP(Reference Lee, Graeff and Walseth39).

CD38 is the most highly investigated ADP-ribosyl cyclase enzyme, and for years it has generated discussion related to its ‘topological paradox’. This paradox questions how CD38, which has an active site facing the exterior of the cell, can regulate the synthesis of an intracellular signalling molecule(Reference De Flora, Guida, Franco and Zocchi104). BST-1 (CD157) shows a similar extracellular location(Reference Itoh, Ishihara, Tomizawa, Tanaka, Kobune, Ishikawa, Kaisho and Hirano101). As explanation, it has been demonstrated that cells possess connexin 43 hemichannels that allow passage of NAD+ from the inside to the outside of the cell(Reference Bruzzone, Guida, Zocchi, Franco and De Flora105). There is also bidirectional transport of cADPR through CD38 itself(Reference Franco, Guida, Bruzzone, Zocchi, Usai and De Flora106). Intracellular NAD+, which is found at micromolar concentrations (as compared with nanomolar concentrations extracellularly), can move down its concentration gradient through the connexon 43 channels to the ectocellular active site of CD38. CD38 can then catalyse the formation of cADPR, which passes through the central channel formed by the homodimeric structure of the protein(Reference De Flora, Zocchi, Guida, Franco and Bruzzone107). Alternatively, cADPR can pass into the cell through nucleoside transporters(Reference Guida, Bruzzone, Sturla, Franco, Zocchi and De Flora108). The process of nucleotide transport can occur via an autocrine mechanism, with the NAD+ and cADPR affecting the emitting cell, or a paracrine mechanism, with the nucleotides affecting cells in the vicinity of the emitting cell(Reference De Flora, Zocchi, Guida, Franco and Bruzzone107). For example, increasing extracellular cADPR increases proliferation in human haematopoietic cells(Reference Podestà, Zocchi and Pitto109) and in 3T3 fibroblasts(Reference Franco, Zocchi, Usai, Guida, Bruzzone, Costa and De Flora110). In the brain, astrocytes respond to extracellular cADPR by increasing intracellular Ca levels which in turn increase neurotransmitter release(Reference Verderio, Bruzzone, Zocchi, Fedele, Schenk, De Flora and Matteoli111), while in bovine tracheal smooth muscle cells, extracellular cADPR increases intracellular Ca and potentiates acetylcholine-induced contraction(Reference Franco, Bruzzone, Song, Guida, Zocchi, Walseth, Crimi, Usai, De Flora and Brusasco112). However, CD38 is also localised to intracellular membranes, including the nucleus and the endoplasmic reticulum(Reference Khoo and Chang97, Reference Adebanjo, Anandatheerthavarada and Koval98, Reference Sun, Adebanjo and Koval113Reference Ceni, Pochon and Brun115), which suggests that this enzyme also has an intracellular site of action. And, as previously mentioned, other soluble and membrane-bound ADP-ribosyl cyclase enzymes have been identified, so there is evidence for both intracellular and extracellular cyclases, although much remains to be understood about their precise roles in cADPR synthesis. Both CD38-(Reference Ceni, Pochon and Brun115) and non-CD38(Reference Ceni, Muller-Steffner, Lund, Pochon, Schweitzer, De Waard, Schuber, Villaz and Moutin43, Reference Ceni, Pochon, Villaz, Muller-Steffner, Schuber, Baratier, De Waard, Ronjat and Moutin103)-dependent ADP-ribosyl cyclase activity has been found in the brain. Distribution of CD38 in both rat(Reference Yamada, Mizuguchi, Otsuka, Ikeda and Takahashi116) and human(Reference Mizuguchi, Otsuka, Sato, Ishii, Kon, Yamada, Nishina, Katada and Ikeda117) brain is widespread.

Cyclic adenosine diphosphate ribose and hippocampal synaptic plasticity

LTP and LTD, which are long-lasting increases and decreases (respectively) in synaptic strength, are used experimentally to model learning and memory(Reference Malenka and Bear118). While there are various forms of LTD and LTP that differ in many respects, in all cases there is an increase in intracellular Ca levels. Induction of LTP requires a substantial rise in intracellular Ca, while a more moderate rise in intracellular Ca results in induction of LTD(Reference Franks and Sejnowski50). At least in N-methyl-d-aspartate acid receptor (NMDAR)-dependent forms, the signal cascade generated following LTP induction involves activation of Ca-dependent protein kinases such as Ca calmodulin kinase II(Reference Malinow, Schulman and Tsien119, Reference Silva, Wang, Paylor, Wehner, Stevens and Tonegawa120), while that generated following LTD induction involves activation of Ca-dependent phosphatases such as calcineurin(Reference Mulkey, Endo, Shenolikar and Malenka121, Reference Mulkey, Herron and Malenka122). With respect to Ca-release channels, RyR are particularly concentrated in the dendritic spines of the hippocampus, in contrast to IP3 receptors, which are concentrated in the dendritic shafts(Reference Sharp, McPherson, Dawson, Aoki, Campbell and Snyder41). The Ca in dendritic spines has been proposed as being especially important in synaptic plasticity(Reference Sabatini, Oertner and Svoboda123), so Ca released from RyR might be particularly essential for modulating hippocampal synaptic function.

There is considerable evidence that cADPR is required for a form of LTD in hippocampal neurons, although the exact mechanism by which cADPR exerts this effect is unclear. Both NMDAR-dependent and metabotropic-glutamate-receptor-dependent forms of LTD are found in the hippocampus of juvenile rats(Reference Nicoll, Oliet and Malenka124), and there is evidence for Ca release from ryanodine-sensitive stores in both. Early studies found that administration of dantrolene, a ryanodine channel blocker, blocked LTD and enhanced LTP in NMDAR-dependent hippocampal LTD(Reference O'Mara, Rowan and Anwyl125), while Ca influx through low-voltage-activated Ca channels and release of Ca from ryanodine-sensitive Ca stores was linked to a form of NMDAR-independent hippocampal LTD(Reference Wang, Rowan and Anwyl126). Later studies suggested that hippocampal LTD induction required release of Ca from both pre- and postsynaptic stores, with a ryanodine-sensitive channel as the presynaptic store and probably IP3 as the postsynaptic store(Reference Reyes and Stanton127). Further investigation of the presynaptic role of RyR in hippocampal LTD determined that NMDAR-dependent LTD is followed by postsynaptic synthesis of NO and presynaptic activation of guanylyl cyclase, which probably enhances cADPR formation and the release of Ca from ryanodine-sensitive stores(Reference Reyes-Harde, Potter, Galione and Stanton10, Reference Reyes-Harde, Empson, Potter, Galione and Stanton11). cADPR was finally shown experimentally to be associated with this presynaptic form of LTD in Reyes-Harde et al. (Reference Reyes-Harde, Empson, Potter, Galione and Stanton11). Presynaptic modulation of LTD involves changes in neurotransmitter release, either through reductions in quantal size or frequency of transmission(Reference Nicoll, Oliet and Malenka124). It was recently shown that the NO/LTD cascade at the frog neuromuscular junction involves activation of calmodulin and the Ca-sensitive enzyme calcineurin(Reference Etherington and Everett128). The authors proposed that in this pathway, there is a long-lasting depression of transmitter release due to sustained activity of the NO signalling pathway following calcineurin-mediated dephosphorylation of NOS, which results in sustained NO production(Reference Etherington and Everett128). As previously discussed, both NMDAR-dependent and metabotropic-glutamate-receptor-dependent forms of LTD could involve modulation of the presynaptic neuron by a retrograde messenger, so these results are consistent with known LTD characteristics.

As just mentioned, release of Ca from ryanodine-sensitive stores was associated with depotentiation of previously established LTP in the rat dentate gyrus(Reference O'Mara, Rowan and Anwyl125). RyR have also been associated with the induction of the late form of LTP in the hippocampus in a process requiring NO, cGMP and cGMP protein-dependent kinase(Reference Lu, Kandel and Hawkins129). As cADPR synthesis is stimulated by cGMP(Reference Galione, White, Willmott, Turner, Potter and Watson54), it may also be involved in this cascade. RyR have been further implicated in weak LTP, as induced by a small conditioning stimulus, but not in the LTP that follows a moderate or strong conditioning protocol(Reference Raymond and Redman130). However, the evidence linking cADPR with LTP is not as clear as for LTD. In the only direct investigation of cADPR and LTP, administration of the cADPR antagonist 8-Br-cADPR had no effect on LTP induction(Reference Reyes-Harde, Empson, Potter, Galione and Stanton11) using an NMDAR-induction protocol. However, there are different types of LTP and several induction protocols that can be used, so this result is not conclusive. Investigations of type III RyR knockout mice, which might provide indirect evidence of the role of cADPR in LTP, have yielded conflicting results. In one study, RyR III knockout mice were found to exhibit decreased LTP in the CA1 region of the hippocampus and a decrease in α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor response, although this did not appear to be due to changes in receptor density(Reference Shimuta, Yoshikawa, Fukaya, Watanabe, Takeshima and Manabe131). In a second study, RyR III knockout mice were found to exhibit facilitated LTP in the CA1 region of the hippocampus, with a corresponding impairment of LTD(Reference Futatsugi, Kato, Ogura, Li, Nagata, Kuwajima, Tanaka, Itohara and Mikoshiba132). And in a third study, RyR III knockout mice showed no change in CA1 LTP(Reference Balschun, Wolfer, Bertocchini, Barone, Conti, Zuschratter, Missiaen, Lipp, Frey and Sorrentino133). Each of these studies used different LTP induction protocols and experimental conditions, which shows that it is difficult to compare LTP results across studies when the same procedures are not used. Also, different strains of mice were used in these experiments, so there may have been differences between knockout models as well.

Due to the importance of the hippocampus in spatial learning(Reference Redish and Touretzky12, Reference Morris, Garrud, Rawlins and O'Keefe134), altered hippocampal synaptic plasticity might be expected to affect this ability. Although there are no studies which directly link cADPR with spatial learning ability, two studies of the RyR III knockout mouse also looked at performance of these animals in the Morris water maze (MWM), which was introduced in 1981 as a tool to investigate spatial learning and memory in neurobehavioural research(Reference Morris135). As with LTP, these studies report different effects of RyR III gene deletion on behaviour. In Futatsugi et al. (Reference Futatsugi, Kato, Ogura, Li, Nagata, Kuwajima, Tanaka, Itohara and Mikoshiba132), knockout mice showed improved spatial learning ability as evidenced by greater spatial accuracy in a probe trial, while in Balschun et al. (Reference Balschun, Wolfer, Bertocchini, Barone, Conti, Zuschratter, Missiaen, Lipp, Frey and Sorrentino133), loss of RyR III had a negative effect on spatial learning ability, with animals showing reduced flexibility in relearning a new platform location. Although these results are not consistent, the studies also found that hippocampal neurons had different electrophysiological properties. The effect of RyR and cADPR on spatial learning ability is not clear at this time, although it has been shown that spatial learning increases the expression of RyR type II in the rat hippocampus(Reference Zhao, Meiri, Xu, Cavallaro, Quattrone, Zhang and Alkon136).

Dietary niacin, brain cyclic adenosine diphosphate ribose and spatial learning

We investigated the effect of dietary niacin on brain cADPR and MWM performance using three different models of niacin deficiency and one model of niacin supplementation(Reference Young, Jacobson and Kirkland137). In each, male weanling Long–Evans rats were used since performance of female rats in the water maze has also been show to vary across the oestrous cycle(Reference Warren and Juraska138) and tryptophan metabolism in females has been observed to change with hormonal variations(Reference Shibata and Kondo139). Long–Evans rats are the most commonly used rats in water maze experiments, and their ability to perform successfully has been well validated(Reference D'Hooge and De Deyn140). Although the use of weanling rats introduces some concerns about potential dietary effects on synaptogenesis and myelination, which are not complete until 60 d after birth(Reference Akiyama, Ichinose, Omori, Sakurai and Asou141), we have previously shown that young rats show a much greater sensitivity to niacin deficiency than older rats, possibly due to a reduced tryptophan →  NAD+ conversion ability (JB Kirkland, unpublished results). The diets used in these experiments are modelled after AIN-93G (designed by the American Institute of Nutrition, formulated to meet gestation, lactation and growth requirements)(Reference Reeves, Nielsen and Fahey142), with 20 % casein replaced by 7 % casein and 6 % gelatin (Table 3). This represents a low level of protein, and tryptophan content is limiting in order to minimise tryptophan →  NAD+ conversion. The micronutrient levels are as described for AIN-93G, with the exception of niacin content in deficient and high-dose diets. Nicotinamide was chosen as the supplemented form of niacin because the brain shows a preference for using nicotinamide in the synthesis of NAD+ over any other precursors, and there is an active mechanism for nicotinamide uptake into the brain, where it is distributed evenly(Reference Spector143). Pharmacological nicotinamide supplementation has been investigated as a treatment for type 1 diabetes in children, so the level of nicotinamide used was comparable with the human consumption of 1–3 g nicotinamide per d in the diabetes prevention trials(Reference Pozzilli, Browne and Kolb144).

Table 3 Composition of experimental diets (g/kg diet)

* Vitamin mix composition: sucrose, 97 543 mg/kg; vitamin B12, 1 mg/kg; vitamin E (dl-α-tocopheryl acetate), 20 000 mg/kg; biotin, 20 mg/kg; calcium pantothenate, 1600 mg/kg; folic acid, 200 mg/kg; vitamin K (phylloquinone), 50 mg/kg; pyridoxine HCl, 700 mg/kg; riboflavin, 600 mg/kg; thiamin HCl, 600 mg/kg; retinyl palmitate, 800 mg/kg; cholecalciferol, 2·5 mg/kg.

For all of our experiments, statistical analyses were performed using SPSS (version 12.0 for Windows; SPSS Inc., Chicago, IL, USA). The P value was set at ≤ 0·05. A trend was defined as a P value between 0·05 and 0·1. The Kolmogorov–Smirnov and Shapiro–Wilk tests were used to evaluate normality. As the water maze data in each experiment showed an abnormal distribution on at least one experimental day, non-parametric tests were used to assess water maze performance. The Friedman test was performed to determine the within-subjects effect, and the Kruskal–Wallis test was performed to determine the between-subjects effect. The within-subjects factor was time (test day) and the between-subjects factor was diet. For probe trial analysis, one-way ANOVA were run comparing the number of platform crossings at each of the four possible platform locations. Two-tailed independent t tests were used to compare mean swim speeds and brain nucleotides. Data were not transformed before analysis.

In the first niacin-deficiency model, niacin-deficient rats were compared with pair-fed controls (n 8). Control rats were pair fed a diet containing 30 mg added nicotinic acid per kg diet throughout the duration of the experiment. This level is considered adequate to fully meet the needs of rats, and is found in AIN-93 formulations and most commercial rat chows. In the water maze, niacin-deficient rats showed superior spatial learning ability during acquisition on day 2 (P = 0·01), day 3 (P = 0·05), day 5 (P = 0·007) and day 6 (P = 0·001) out of 7 d of testing, and tended to do so on day 4 (P = 0·1) (Fig. 3(a)). There was also a trend (P = 0·09) for higher spatial accuracy in a probe test. Brain NAD+ was decreased by 42 % and brain cADPR by 36 % (Table 4). The pair-feeding model was used to control for differences in feed intake between niacin-deficient and control rats, since a deficiency of niacin, like most micronutrients, causes anorexia(Reference Kirkland, Rawling, Rucker, Zempleni, Suttie and McCormick145). In our experiments, niacin-deficient weanling rats usually consume between 5 and 8 g food per d over periods of up to 5 weeks(Reference Young, Jacobson and Kirkland137). In contrast, food intake of healthy rats should increase during the growth period, and normal levels can be more than twice that consumed during niacin deficiency(Reference Ahmed, Bedi, Warren and Kamel146). Both niacin-deficient and pair-fed rats are consequently significantly deprived of food, particularly in the later experimental weeks, and show a significantly reduced body weight when compared with normative rat growth charts.

Fig. 3 (a) Cumulative error of niacin-deficient (–●–) and pair-fed (–○–) rats in the water maze. Rats were tested in three daily trials across 6 d with an inter-trial interval of 2 h. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means (n 8), with their standard errors represented by vertical bars. * Mean value was significantly different from that of the pair-fed rats (P ≤ 0·05). (b) Cumulative error of niacin-deficient (–●–; n 9) and partially feed-restricted (–○–; n 8) rats in the water maze. Rats were tested in three daily trials across 6 d with an inter-trial interval of 2 h. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means, with their standard errors represented by vertical bars. * Mean value was significantly different from that of the partially feed-restricted rats (P ≤ 0·05). (c) Cumulative error of niacin-deficient (–●–) and niacin-recovered (–○–) rats during reversal training in the water maze. Rats were tested in three daily trials across 4 d with an inter-trial interval of 2 h. The reversal training followed an initial acquisition phase in the water maze and 4 d of niacin refeeding. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means (n 9), with their standard errors represented by vertical bars. * Mean value was significantly different from that of the niacin-recovered rats (P ≤ 0·05). (d) Cumulative error of niacin-supplemented (–●–; n 18) and control (–○–; n 15) rats in the water maze. Rats were tested in three daily trials across 6 d with an inter-trial interval of 2 h. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means, with their standard errors represented by vertical bars. * Mean value was significantly different from that of the control rats (P ≤ 0·05). (e) Proximity averages to the platform during hidden platform testing by Cd38− / −  (–●–) and wild-type (–○–) mice across 7 d of testing. Mice were tested in three daily trials across 6 d with an inter-trial interval of 2 h. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means (n 10), with their standard errors represented by vertical bars. * Mean value was significantly different from that of the wild-type rats (P ≤ 0·05). Fig. 3(a–d) were originally published in Young et al. (2007)(Reference Young, Jacobson and Kirkland137). Fig. 3(e) was originally published in Young et al. (2008)(Reference Young, Choleris, Lund and Kirkland149).

Table 4 Brain NAD+ and cyclic adenosine diphosphate ribose (cADPR) in rats with differing niacin intakes and in Cd38−/− mice (nmol/g tissue) (Mean values with their standard errors)

* Mean value was significantly different from that of rats in the comparative group (P ≤ 0·05).

In the second niacin-deficiency model, niacin-deficient rats were compared with partially feed-restricted controls (nine niacin-deficient rats and eight partially feed-restricted rats). Control rats were pair fed for the first 16 d of the experiment, and then fed ad libitum for 4 d before and during the entire period of water maze testing. In the water maze, we observed that niacin-deficient rats again showed superior performance on day 3 and day 4 (P < 0·05) of 6 d of testing, and tended to do so on day 6 (P = 0·08) (Fig. 3(b)), although the spatial accuracy of the two groups was comparable in a probe test. Brain NAD+ and cADPR were not measured. The food intake and body weight of the two groups diverged greatly once the control group was placed on ad libitum feeding. The goal of this feeding strategy was to reduce hunger during the period of behavioural testing, while minimising the developmental differences that would result between niacin-deficient rats and rats fed ad libitum throughout the entire experiment.

In the third niacin-deficiency model, currently niacin-deficient rats were compared with niacin-recovered rats (n 9). All rats were maintained on a niacin-deficient diet during the first phase of water maze testing, and then half were recovered from the deficiency through niacin refeeding during the second phase of water maze testing. Nicotinamide was used since it is the preferred substrate for NAD+ in most tissues, including the brain(Reference Jacobson, Dame, Pyrek and Jacobson147) and should therefore allow for a more rapid replenishment of niacin metabolites. The 4 d period of niacin refeeding before the second phase of water maze testing was designed to mimic the period of nutritional rehabilitation typically required for resolution of the symptoms of pellagrous dementia(Reference Kirkland, Rawling, Rucker, Zempleni, Suttie and McCormick145). During the first phase of water maze testing, when all rats were niacin deficient, a retrospective analysis of performance revealed no significant differences between the two groups. However, during the second phase of water maze testing, when the recovered rats were being refed niacin, there was a significant effect of diet on day 4 (P = 0·01) and a trend on day 2 (P = 0·09) (Fig. 3(c)), with comparable spatial accuracy in the probe test. It is important to note that during the second phase of testing, rats were no longer naive and had already undergone extensive water maze training, so the sensitivity of the test to detect subtle differences in spatial learning ability would be reduced. Brain NAD+ was decreased by 43 % and brain cADPR by 25 % (Table 4). This approach sought to determine the flexibility of learning and brain cADPR to niacin refeeding.

In the niacin-supplementation model, niacin-supplemented rats were compared with control rats (eighteen supplemented rats and fifteen control rats). Niacin-supplemented rats were fed a diet containing 4 g added nicotinamide/kg diet. The amount of food eaten daily during this period by each rat was determined and was correlated to body weight, allowing for an estimation of the average amount of food eaten daily per g body weight. The amount was initially determined at 0·15 g food/g body weight, and this value was used to calculate the amount of food provided on each day of the experimental period. This level of food provision was used throughout the experiment, as the daily residual food suggested that it allowed ad libitum, or nearly ad libitum, feeding to all animals. In the water maze, supplemented rats showed inferior spatial learning ability on day 3 (P = 0·04) of 6 d of testing (Fig. 3(d)). Brain NAD+ was increased by 38 % and brain cADPR by 14 % (Table 4).

The consistency of the finding of improved spatial learning ability in each niacin-deficiency model is striking. This, combined with the opposite observation following nicotinamide supplementation, is supportive of an inverse relationship between spatial learning ability and dietary niacin intake in very young rats. Although the link between dietary niacin, spatial learning ability and cADPR is correlational, dietary niacin might affect brain function through cADPR modulation. Like niacin, brain cADPR shows an inverse relationship with spatial learning ability, and while cADPR levels are quickly restored to normal following niacin refeeding, the cognitive benefits associated with the deficiency rapidly disappear.

CD38, brain cyclic adenosine diphosphate ribose and spatial learning

We also investigated the effect of CD38 gene deletion on brain cADPR(Reference Young, Choleris, Lund and Kirkland148) and MWM performance(Reference Young, Choleris, Lund and Kirkland149). The Cd38− / −  mouse was originally generated to study the role of CD38 in humoral immunity, and has subsequently been used in investigations of airway smooth muscle function(Reference Deshpande, White, Guedes, Milla, Walseth, Lund and Kannan150), glucose tolerance(Reference Kato, Yamamoto, Fujimura, Noguchi, Takasawa and Okamoto151), osteogenesis(Reference Sun, Iqbal and Dolgilevich152) and innate immunity(Reference Partida-Sánchez, Cockayne and Monard56, Reference Partida-Sanchez, Randall and Lund153). While the original study of the Cd38− / −  mouse reported comparable behaviour between knockout and wild-type animals(Reference Cockayne, Muchamuel, Grimaldi, Muller-Steffner, Randall, Lund, Murray, Schuber and Howard154), there have been no comprehensive investigations of specific behavioural functions in this transgenic model. Cd38− / −  mice were generated by gene targeting(Reference Cockayne, Muchamuel, Grimaldi, Muller-Steffner, Randall, Lund, Murray, Schuber and Howard154) and were backcrossed for twelve generations to C57BL/6J(Reference Partida-Sánchez, Cockayne and Monard56). This practice complies with the recommendation of the Banbury Conference that targeted mutations be maintained in congenic lines(155), although it is nonetheless likely that the knockouts carry alleles for genes that flank the mutation locus(Reference Crusio156). Backcrossing for twelve generations would reduce the length of the chromosome segment from the background genotype to about 16 cM, which when considered in relation to the mouse genome, would contain approximately 300 genes(Reference Gerlai157). So, the Cd38− / −  mouse would contain more than 99 % C57BL/6J genes. When inbred mouse strains are evaluated in the MWM, C57BL/6 mice are often characterised as being the strain of choice, and their ability to learn the task has been validated experimentally(Reference Crawley, Belknap and Collins158).

We observed that brain cADPR was increased (P < 0·001) in the Cd38− / −  mouse as compared with wild-type controls (n 15)(Reference Young, Choleris, Lund and Kirkland148) (Table 4). This is in contrast to Partida-Sánchez et al. (Reference Partida-Sánchez, Cockayne and Monard56) and Ceni et al. (Reference Ceni, Pochon, Villaz, Muller-Steffner, Schuber, Baratier, De Waard, Ronjat and Moutin103), who previously measured levels of cADPR in these tissues and found them to be non-significantly decreased. While our levels of brain cADPR are comparable with other published reports, the degree of variability in each group is reduced, which we believe is due to modifications that we have made to the fluorimetric cycling assay for cADPR. These modifications are shown to increase the recovery of cADPR, improve the functionality of the assay, and reduce between-subject variability(Reference Young and Kirkland159). In fact, we observed a significant reduction of brain cADPR despite a difference of only 16% between wild-type and knockout mice, in contrast to the 20% non-significant reduction observed by Partida-Sanchez et al. (Reference Partida-Sánchez, Cockayne and Monard56) and the 18% non-significant reduction observed in Ceni et al. (Reference Ceni, Pochon, Villaz, Muller-Steffner, Schuber, Baratier, De Waard, Ronjat and Moutin103). We also observed that levels of NAD+ in the brain were increased by 160%(Reference Young, Choleris, Lund and Kirkland148) (P < 0·001) (Table 4). CD38 is a multifunctional enzyme that functions as both a cyclase and a hydrolase enzyme, forming ADP-ribose from hydrolysis of NAD+ or cADPR. Since the ratio of cyclase:hydrolase activity is low(Reference Schuber and Lund160), the loss of NAD+ hydrolase activity might explain an increase in NAD+ of this magnitude, and the differential activity of this enzyme would result in a much greater effect on NAD+ increase than on cADPR reduction, which is what we observed.

We also observed that like niacin-deficient rats, Cd38− / −  mice show improved performance in the MWM as compared with wild-type controls (n 10). Cd38− / −  mice had a significantly shorter latency to the hidden platform on day 5 (P = 0·05) of 7 d of testing, and there was a trend for a shorter latency on day 7 (P = 0·1). Analysis of the proximity average, which takes into account how close the animal comes to the platform(Reference Gallagher, Burwell and Burchinal161), confirmed that on day 5 Cd38− / −  mice performed significantly better than wild-type mice (P = 0·001), while on day 7, there was a trend (P = 0·07) for better performance by the Cd38− / −  mice (Fig. 3(e)). The mean proximity averages were lower for Cd38− / −  mice from day 4 to day 7, demonstrating a consistent pattern for Cd38− / −  mice to perform better than wild-type mice in the water maze on these days. In the probe trial, there was a trend (P = 0·07) for Cd38− / −  mice to cross the target location more times than wild-type mice. Although the effect of CD38 gene deletion on water maze performance was less than seen in niacin deficiency, the results are nonetheless consistent with our previous observation of reduced brain cADPR and improved spatial learning ability in niacin-deficient rats. The magnitude of cADPR change was greater in the niacin-deficiency models than in Cd38− / −  mice (25–35 v. 16 %), so when considered relative to this, these results suggest that although CD38 forms only a proportion of brain cADPR, its removal impacts on brain function by a similar mechanism as that of niacin deficiency. Unlike niacin-deficient rats, which have reduced brain NAD+(Reference Young, Jacobson and Kirkland137), the spatial learning effect in Cd38− / −  mice is observed with increased brain NAD+.

This was a revealing observation, as it identified total cADPr, and/or CD38-catalytic activity (capable of cADPr or NAADP synthesis), as parameters that correlated with MWM performance across all models (niacin-deficient/control/pharmacological nicotinic acid diets in rats, Cd38− / −  v. wild-type mice). Conversely, brain NAD+, and, by extension, the assumed activity of all other NAD+-dependent enzymes did not correlate with performance across all models.

Conclusion

The link between niacin and cADPR provides a fresh insight into the pathophysiology of pellagra, particularly with respect to pellagrous dementia, for which explanations based on redox reactions and energy metabolism did not adequately explain the aetiology of the clinical symptoms. Recovery of psychological function within days when insane pellagra patients are treated with niacin(Reference Spies, Bean, Ashe and Carpenter7) suggests that the dementia is caused by alterations in neural signalling pathways, rather than structural pathological changes. Altered Ca signalling due to changes in levels of cADPR could provide the missing link between the vitamin deficiency and the symptoms of the disease. Our findings of enhanced spatial learning in niacin-deficient rats and in Cd38− / −  mice may be due to changes in the hippocampal synaptic plasticity, and there is some published evidence to support this hypothesis. In a knockout model, deletion of type III RyR improved spatial learning ability, impaired LTD, and facilitated LTP(Reference Takeshima, Ikemoto and Nishi162). Although cADPR was not directly implicated in this study, cADPR binds to RyR III(Reference Higashida, Hashii, Yokoyama, Hoshi, Chen, Egorova, Noda and Zhang163), so the effects of gene deletion may be due to reduced cADPR-induced Ca signalling, which would also be observed with niacin deficiency. However, in previous work, loss of RyR III had a negative effect on spatial learning ability, with animals showing reduced flexibility in relearning a new platform location(Reference Balschun, Wolfer, Bertocchini, Barone, Conti, Zuschratter, Missiaen, Lipp, Frey and Sorrentino133), so there are conflicting reports. Other studies have shown that animals which show facilitated LTD display impaired spatial learning(Reference Yang, Han, Cao, Li and Xu164, Reference Xiong, Yang, Cao, Wei, Liang, Yang and Xu165), so if LTD is indeed impaired by decreased cADPR, spatial learning might be expected to improve. Further studies are required to investigate precisely the effects of niacin and cADPR on hippocampal electrophysiology. We are currently exploring this avenue as well as investigating the effect of dietary niacin and water maze training on gene expression in the hippocampus.

Although the evidence linking cADPR, CD38 and spatial learning ability presented in the present review is correlational, consistency of the findings across several different models greatly strengthens this relationship. Further validation comes from the observation that changes in spatial learning ability vary proportionately with the degree of changes in brain cADPR concentration. Niacin-deficient rats show the greatest decrease in cADPR, and the greatest improvement in spatial learning ability, while Cd38− / −  mice show a more modest decrease in cADPR, and a more modest improvement in spatial learning ability. Unlike in niacin-deficient rats, this occurred with an increase in brain NAD+, providing support for the causative role of cADPR in altered maze performance. In contrast, niacin-supplemented rats show a small increase in cADPR, and a small spatial learning impairment. These findings are supportive of an inverse relationship between spatial learning ability and dietary niacin intake in very young rats, although a direct link between cADPR and spatial learning ability is still missing. Clearly, brain cADPR and spatial learning ability are significantly affected by dietary niacin. Intake of this nutrient may therefore play a role in the molecular events regulating learning performance, and further investigations of niacin intake, CD38 and cADPR may help identify potential molecular targets for clinical intervention to enhance learning and prevent or reverse cognitive decline.

Acknowledgements

The present review was supported by funding from the Natural Sciences and Engineering Research Council (NSERC) of Canada and National Institutes of Health (NIH) grant CA-43894. There are no conflicts of interest.

G. S. Y. was the primary author of the paper which was adapted from her PhD thesis. J B K. was G. S. Y.'s advisor.

References

1Kirkland, JB (2006) Niacin. In Handbook of Vitamins, 4th ed., pp. 191232 [Rucker, RB, Zempleni, J, Suttie, JW and McCormick, DB, editors]. Boca Raton, FL: Taylor & Francis.Google Scholar
2Preiss, J & Handler, P (1958) Biosynthesis of diphosphopyridine nucleotide. I. Identification of intermediates. J Biol Chem 233, 488492.CrossRefGoogle ScholarPubMed
3Dietrich, LS, Fuller, L, Yero, IL & Martinez, L (1966) Nicotinamide mononucleotide pyrophosphorylase activity in animal tissues. J Biol Chem 241, 188191.Google Scholar
4Shibata, K, Mushiage, M, Kondo, T, Hayakawa, T & Tsuge, H (1995) Effects of vitamin B6 deficiency on the conversion ratio of tryptophan to niacin. Biosci Biotechnol Biochem 59, 20602063.Google Scholar
5Shibata, K (1999) Nutritional factors that regulate on the conversion of l-tryptophan to niacin. Adv Exp Med Biol 467, 711716.Google Scholar
6Buniva, (1981) Observations in pellagra: it would not appear to be contagious. In Pellagra, pp. 1112 [Carpenter, KJ, editor]. Stroudsburg, PA: Hutchinson Ross.Google Scholar
7Spies, TD, Bean, WB & Ashe, WF (1981) Recent advances in the treatment of pellagra and associated deficiencies. In Pellagra, pp. 213225 [Carpenter, KJ, editor]. Stroudsburg, PA: Hutchinson Ross.Google Scholar
8Hoffer, A (1970) Pellagra and schizophrenia. Psychosomatics 11, 522525.Google Scholar
9Lee, HC, Walseth, TF, Bratt, GT, Hayes, RN & Clapper, DL (1989) Structural determination of a cyclic metabolite of NAD+ with intracellular Ca2+-mobilizing activity. J Biol Chem 264, 16081615.Google Scholar
10Reyes-Harde, M, Potter, BV, Galione, A & Stanton, PK (1999) Induction of hippocampal LTD requires nitric-oxide-stimulated PKG activity and Ca2+ release from cyclic ADP-ribose-sensitive stores. J Neurophysiol 82, 15691576.CrossRefGoogle ScholarPubMed
11Reyes-Harde, M, Empson, R, Potter, BV, Galione, A & Stanton, PK (1999) Evidence of a role for cyclic ADP-ribose in long-term synaptic depression in hippocampus. Proc Natl Acad Sci U S A 96, 40614066.CrossRefGoogle ScholarPubMed
12Redish, AD & Touretzky, DS (1998) The role of the hippocampus in solving the Morris water maze. Neural Comput 10, 73111.CrossRefGoogle ScholarPubMed
13Berridge, MJ, Lipp, P & Bootman, MD (2000) The versatility and universality of calcium signalling. Nat Rev Mol Cell Biol 1, 1121.Google Scholar
14Carafoli, E, Santella, L, Branca, D & Brini, M (2001) Generation, control, and processing of cellular calcium signals. Crit Rev Biochem Mol Biol 36, 107260.Google Scholar
15Meldolesi, J & Pozzan, T (1998) The endoplasmic reticulum Ca2+ store: a view from the lumen. Trends Biochem Sci 23, 1014.CrossRefGoogle Scholar
16Pozzan, T, Magalhaes, P & Rizzuto, R (2000) The comeback of mitochondria to calcium signalling. Cell Calcium 28, 279283.CrossRefGoogle ScholarPubMed
17Gilabert, JA, Bakowski, D & Parekh, AB (2001) Energized mitochondria increase the dynamic range over which inositol 1,4,5-trisphosphate activates store-operated calcium influx. EMBO J 20, 26722679.CrossRefGoogle ScholarPubMed
18Collins, TJ, Berridge, MJ, Lipp, P & Bootman, MD (2002) Mitochondria are morphologically and functionally heterogeneous within cells. EMBO J 21, 16161627.CrossRefGoogle ScholarPubMed
19Malviya, AN, Rogue, P & Vincendon, G (1990) Stereospecific inositol 1,4,5-[32P]trisphosphate binding to isolated rat liver nuclei: evidence for inositol trisphosphate receptor-mediated calcium release from the nucleus. Proc Natl Acad Sci U S A 87, 92709274.Google Scholar
20Nicotera, P, Orrenius, S, Nilsson, T & Berggren, PO (1990) An inositol 1,4,5-trisphosphate-sensitive Ca2+ pool in liver nuclei. Proc Natl Acad Sci U S A 87, 68586862.Google Scholar
21Gerasimenko, OV, Gerasimenko, JV, Tepikin, AV & Petersen, OH (1995) ATP-dependent accumulation and inositol trisphosphate- or cyclic ADP-ribose-mediated release of Ca2+ from the nuclear envelope. Cell 80, 439444.CrossRefGoogle ScholarPubMed
22Pinton, P, Pozzan, T & Rizzuto, R (1998) The Golgi apparatus is an inositol 1,4,5-trisphosphate-sensitive Ca2+ store, with functional properties distinct from those of the endoplasmic reticulum. EMBO J 17, 52985308.CrossRefGoogle ScholarPubMed
23Yoo, SH (2000) Coupling of the IP3 receptor/Ca2+ channel with Ca2+ storage proteins chromogranins A and B in secretory granules. Trends Neurosci 23, 424428.CrossRefGoogle Scholar
24Gerasimenko, JV, Tepikin, AV, Petersen, OH & Gerasimenko, OV (1998) Calcium uptake via endocytosis with rapid release from acidifying endosomes. Curr Biol 8, 13351338.CrossRefGoogle ScholarPubMed
25Lee, HC & Aarhus, R (1995) A derivative of NADP mobilizes calcium stores insensitive to inositol trisphosphate and cyclic ADP-ribose. J Biol Chem 270, 21522157.Google Scholar
26Streb, H, Irvine, RF, Berridge, MJ & Schulz, I (1983) Release of Ca2+ from a nonmitochondrial intracellular store in pancreatic acinar cells by inositol-1,4,5-trisphosphate. Nature 306, 6769.Google Scholar
27Mikoshiba, K, Furuichi, T, Miyawaki, A, et al. . (1993) Inositol trisphosphate receptor and Ca2+ signalling. Philos Trans R Soc Lond B Biol Sci 340, 345349.Google ScholarPubMed
28Moore-Nichols, D, Arnott, A & Dunn, RC (2002) Regulation of nuclear pore complex conformation by IP(3) receptor activation. Biophys J 83, 14211428.Google Scholar
29Surroca, A & Wolff, D (2000) Inositol 1,4,5-trisphosphate but not ryanodine-receptor agonists induces calcium release from rat liver Golgi apparatus membrane vesicles. J Membr Biol 177, 243249.CrossRefGoogle Scholar
30Gerasimenko, JV, Sherwood, M, Tepikin, AV, Petersen, OH & Gerasimenko, OV (2006) NAADP, cADPR and IP3 all release Ca2+ from the endoplasmic reticulum and an acidic store in the secretory granule area. J Cell Sci 119, 226238.CrossRefGoogle Scholar
31Gerasimenko, O & Gerasimenko, J (2004) New aspects of nuclear calcium signalling. J Cell Sci 117, 30873094.Google Scholar
32Churchill, GC, Okada, Y, Thomas, JM, Genazzani, AA, Patel, S & Galione, A (2002) NAADP mobilizes Ca(2+) from reserve granules, lysosome-related organelles, in sea urchin eggs. Cell 111, 703708.CrossRefGoogle ScholarPubMed
33Gerasimenko, JV, Maruyama, Y, Yano, K, Dolman, NJ, Tepikin, AV, Petersen, OH & Gerasimenko, OV (2003) NAADP mobilizes Ca2+ from a thapsigargin-sensitive store in the nuclear envelope by activating ryanodine receptors. J Cell Biol 163, 271282.CrossRefGoogle ScholarPubMed
34Mikoshiba, K, Furuichi, T & Miyawaki, A (1994) Structure and function of IP3 receptors. Semin Cell Biol 5, 273281.Google Scholar
35Guse, AH (2005) Second messenger function and the structure-activity relationship of cyclic adenosine diphosphoribose (cADPR). FEBS J 272, 45904597.CrossRefGoogle ScholarPubMed
36Galione, A & Petersen, OH (2005) The NAADP receptor: new receptors or new regulation? Mol Interv 5, 7379.CrossRefGoogle ScholarPubMed
37Hohenegger, M, Suko, J, Gscheidlinger, R, Drobny, H & Zidar, A (2002) Nicotinic acid-adenine dinucleotide phosphate activates the skeletal muscle ryanodine receptor. Biochem J 367, 423431.Google Scholar
38Mojzisova, A, Krizanova, O, Zacikova, L, Kominkova, V & Ondrias, K (2001) Effect of nicotinic acid adenine dinucleotide phosphate on ryanodine calcium release channel in heart. Pflugers Arch 441, 674677.Google Scholar
39Lee, HC, Graeff, R & Walseth, TF (1995) Cyclic ADP-ribose and its metabolic enzymes. Biochimie 77, 345355.CrossRefGoogle ScholarPubMed
40Chini, EN & Dousa, TP (1996) Nicotinate-adenine dinucleotide phosphate-induced Ca(2+)-release does not behave as a Ca(2+)-induced Ca(2+)-release system. Biochem J 316, 709711.CrossRefGoogle ScholarPubMed
41Sharp, AH, McPherson, PS, Dawson, TM, Aoki, C, Campbell, KP & Snyder, SH (1993) Differential immunohistochemical localization of inositol 1,4,5-trisphosphate- and ryanodine-sensitive Ca2+ release channels in rat brain. J Neurosci 13, 30513063.Google Scholar
42Patel, S, Churchill, GC, Sharp, T & Galione, A (2000) Widespread distribution of binding sites for the novel Ca2+-mobilizing messenger, nicotinic acid adenine dinucleotide phosphate, in the brain. J Biol Chem 275, 3649536497.Google Scholar
43Ceni, C, Muller-Steffner, H, Lund, F, Pochon, N, Schweitzer, A, De Waard, M, Schuber, F, Villaz, M & Moutin, MJ (2003) Evidence for an intracellular ADP-ribosyl cyclase/NAD+-glycohydrolase in brain from CD38-deficient mice. J Biol Chem 278, 4067040678.CrossRefGoogle ScholarPubMed
44Aarhus, R, Graeff, RM, Dickey, DM, Walseth, TF & Lee, HC (1995) ADP-ribosyl cyclase and CD38 catalyze the synthesis of a calcium-mobilizing metabolite from NADP. J Biol Chem 270, 3032730333.Google Scholar
45da Silva, CP & Guse, AH (2000) Intracellular Ca(2+) release mechanisms: multiple pathways having multiple functions within the same cell type? Biochim Biophys Acta 1498, 122133.CrossRefGoogle ScholarPubMed
46Melendez, AJ & Allen, JM (2002) Phospholipase D and immune receptor signalling. Semin Immunol 14, 4955.CrossRefGoogle ScholarPubMed
47Young, KW & Nahorski, SR (2001) Intracellular sphingosine 1-phosphate production: a novel pathway for Ca2+ release. Semin Cell Dev Biol 12, 1925.Google Scholar
48Lee, HC (2001) Physiological functions of cyclic ADP-ribose and NAADP as Ca messengers. Annu Rev Pharmacol Toxicol 41, 317345.CrossRefGoogle ScholarPubMed
49Berridge, MJ (1998) Neuronal calcium signaling. Neuron 21, 1326.Google Scholar
50Franks, KM & Sejnowski, TJ (2002) Complexity of calcium signaling in synaptic spines. Bioessays 24, 11301144.CrossRefGoogle ScholarPubMed
51Scemes, E (2000) Components of astrocytic intercellular calcium signaling. Mol Neurobiol 22, 167179.Google Scholar
52Guse, AH (2004) Regulation of calcium signaling by the second messenger cyclic adenosine diphosphoribose (cADPR). Curr Mol Med 4, 239248.Google Scholar
53Galione, A, Lee, HC & Busa, WB (1991) Ca(2+)-induced Ca2+ release in sea urchin egg homogenates: modulation by cyclic ADP-ribose. Science 253, 11431146.CrossRefGoogle ScholarPubMed
54Galione, A, White, A, Willmott, N, Turner, M, Potter, BV & Watson, SP (1993) cGMP mobilizes intracellular Ca2+ in sea urchin eggs by stimulating cyclic ADP-ribose synthesis. Nature 365, 456459.CrossRefGoogle ScholarPubMed
55Guse, AH, Berg, I, da Silva, CP, Potter, BV & Mayr, GW (1997) Ca2+ entry induced by cyclic ADP-ribose in intact T-lymphocytes. J Biol Chem 272, 85468550.Google Scholar
56Partida-Sánchez, S, Cockayne, DA, Monard, S, et al. . (2001) Cyclic ADP-ribose production by CD38 regulates intracellular calcium release, extracellular calcium influx and chemotaxis in neutrophils and is required for bacterial clearance in vivo. Nat Med 7, 12091216.Google Scholar
57Meszaros, LG, Bak, J & Chu, A (1993) Cyclic ADP-ribose as an endogenous regulator of the non-skeletal type ryanodine receptor Ca2+ channel. Nature 364, 7679.Google Scholar
58Lukyanenko, V & Gyorke, S (1999) Ca2+ sparks and Ca2+ waves in saponin-permeabilized rat ventricular myocytes. J Physiol 521, 575585.Google Scholar
59Cui, Y, Galione, A & Terrar, DA (1999) Effects of photoreleased cADP-ribose on calcium transients and calcium sparks in myocytes isolated from guinea-pig and rat ventricle. Biochem J 342, 269273.CrossRefGoogle ScholarPubMed
60Kuemmerle, JF & Makhlouf, GM (1995) Agonist-stimulated cyclic ADP ribose. Endogenous modulator of Ca(2+)-induced Ca2+ release in intestinal longitudinal muscle. J Biol Chem 270, 2548825494.Google Scholar
61Wilson, HL, Dipp, M, Thomas, JM, Lad, C, Galione, A & Evans, AM (2001) ADP-ribosyl cyclase and cyclic ADP-ribose hydrolase act as a redox sensor. A primary role for cyclic ADP-ribose in hypoxic pulmonary vasoconstriction. J Biol Chem 276, 1118011188.CrossRefGoogle ScholarPubMed
62Kannan, MS, Fenton, AM, Prakash, YS & Sieck, GC (1996) Cyclic ADP-ribose stimulates sarcoplasmic reticulum calcium release in porcine coronary artery smooth muscle. Am J Physiol 270, H801H806.Google ScholarPubMed
63Prakash, YS, Kannan, MS, Walseth, TF & Sieck, GC (1998) Role of cyclic ADP-ribose in the regulation of [Ca2+]i in porcine tracheal smooth muscle. Am J Physiol 274, C1653C1660.Google Scholar
64Li, PL, Tang, WX, Valdivia, HH, Zou, AP & Campbell, WB (2001) cADP-ribose activates reconstituted ryanodine receptors from coronary arterial smooth muscle. Am J Physiol Heart Circ Physiol 280, H208H215.Google Scholar
65Li, PL, Zou, AP & Campbell, WB (1998) Regulation of KCa-channel activity by cyclic ADP-ribose and ADP-ribose in coronary arterial smooth muscle. Am J Physiol 275, H1002H1010.Google ScholarPubMed
66Tang, WX, Chen, YF, Zou, AP, Campbell, WB & Li, PL (2002) Role of FKBP12·6 in cADPR-induced activation of reconstituted ryanodine receptors from arterial smooth muscle. Am J Physiol Heart Circ Physiol 282, H1304H1310.Google Scholar
67Boittin, FX, Dipp, M, Kinnear, NP, Galione, A & Evans, AM (2003) Vasodilation by the calcium-mobilizing messenger cyclic ADP-ribose. J Biol Chem 278, 96029608.Google Scholar
68Sitsapesan, R & Williams, AJ (1995) Cyclic ADP-ribose and related compounds activate sheep skeletal sarcoplasmic reticulum Ca2+ release channel. Am J Physiol 268, C1235C1240.Google Scholar
69Lopez, JR, Cordovez, G, Linares, N & Allen, PD (2000) Cyclic ADP-ribose induces a larger than normal calcium release in malignant hyperthermia-susceptible skeletal muscle fibers. Pflugers Arch 440, 236242.Google Scholar
70Fulceri, R, Rossi, R, Bottinelli, R, Conti, A, Intravaia, E, Galione, A, Benedetti, A, Sorrentino, V & Reggiani, C (2001) Ca2+ release induced by cyclic ADP ribose in mice lacking type 3 ryanodine receptor. Biochem Biophys Res Commun 288, 697702.CrossRefGoogle ScholarPubMed
71Koshiyama, H, Lee, HC & Tashjian, AH Jr (1991) Novel mechanism of intracellular calcium release in pituitary cells. J Biol Chem 266, 1698516988.Google Scholar
72Currie, KP, Swann, K, Galione, A & Scott, RH (1992) Activation of Ca(2+)-dependent currents in cultured rat dorsal root ganglion neurones by a sperm factor and cyclic ADP-ribose. Mol Biol Cell 3, 14151425.Google Scholar
73Morikawa, H, Khodakhah, K & Williams, JT (2003) Two intracellular pathways mediate metabotropic glutamate receptor-induced Ca2+ mobilization in dopamine neurons. J Neurosci 23, 149157.Google Scholar
74White, AM, Watson, SP & Galione, A (1993) Cyclic ADP-ribose-induced Ca2+ release from rat brain microsomes. FEBS Lett 318, 259263.CrossRefGoogle ScholarPubMed
75Budde, T, Sieg, F, Braunewell, KH, Gundelfinger, ED & Pape, HC (2000) Ca2+-induced Ca2+ release supports the relay mode of activity in thalamocortical cells. Neuron 26, 483492.Google Scholar
76Shirasaki, T, Houtani, T, Sugimoto, T & Matsuda, H (2001) Spontaneous transient outward currents: modulation by nociceptin in murine dentate gyrus granule cells. Brain Res 917, 191205.CrossRefGoogle ScholarPubMed
77Morita, K, Kitayama, S & Dohi, T (1997) Stimulation of cyclic ADP-ribose synthesis by acetylcholine and its role in catecholamine release in bovine adrenal chromaffin cells. J Biol Chem 272, 2100221009.Google Scholar
78Mothet, JP, Fossier, P, Meunier, FM, Stinnakre, J, Tauc, L & Baux, G (1998) Cyclic ADP-ribose and calcium-induced calcium release regulate neurotransmitter release at a cholinergic synapse of Aplysia. J Physiol 507, 405414.CrossRefGoogle Scholar
79Bourguignon, LY, Chu, A, Jin, H & Brandt, NR (1995) Ryanodine receptor-ankyrin interaction regulates internal Ca2+ release in mouse T-lymphoma cells. J Biol Chem 270, 1791717922.CrossRefGoogle ScholarPubMed
80Guse, AH, da Silva, CP, Emmrich, F, Ashamu, GA, Potter, BV & Mayr, GW (1995) Characterization of cyclic adenosine diphosphate-ribose-induced Ca2+ release in T lymphocyte cell lines. J Immunol 155, 33533359.Google Scholar
81Inngjerdingen, M, Al Aoukaty, A, Damaj, B & Maghazachi, AA (1999) Differential utilization of cyclic ADP-ribose pathway by chemokines to induce the mobilization of intracellular calcium in NK cells. Biochem Biophys Res Commun 262, 467472.Google Scholar
82Thorn, P, Gerasimenko, O & Petersen, OH (1994) Cyclic ADP-ribose regulation of ryanodine receptors involved in agonist evoked cytosolic Ca2+ oscillations in pancreatic acinar cells. EMBO J 13, 20382043.Google Scholar
83Gobel, A, Krause, E, Feick, P & Schulz, I (2001) IP(3) and cyclic ADP-ribose induced Ca(2+) release from intracellular stores of pancreatic acinar cells from rat in primary culture. Cell Calcium 29, 2937.Google Scholar
84Noguchi, N, Takasawa, S, Nata, K, Tohgo, A, Kato, I, Ikehata, F, Yonekura, H & Okamoto, H (1997) Cyclic ADP-ribose binds to FK506-binding protein 12·6 to release Ca2+ from islet microsomes. J Biol Chem 272, 31333136.Google Scholar
85Takasawa, S, Nata, K, Yonekura, H & Okamoto, H (1993) Cyclic ADP-ribose in insulin secretion from pancreatic β cells. Science 259, 370373.Google Scholar
86Gromada, J, Jorgensen, TD & Dissing, S (1995) Cyclic ADP-ribose and inositol 1,4,5-triphosphate mobilizes Ca2+ from distinct intracellular pools in permeabilized lacrimal acinar cells. FEBS Lett 360, 303306.Google Scholar
87Ozawa, T & Nishiyama, A (1997) Characterization of ryanodine-sensitive Ca2+ release from microsomal vesicles of rat parotid acinar cells: regulation by cyclic ADP-ribose. J Membr Biol 156, 231239.Google Scholar
88Rusinko, N & Lee, HC (1989) Widespread occurrence in animal tissues of an enzyme catalyzing the conversion of NAD+ into a cyclic metabolite with intracellular Ca2+-mobilizing activity. J Biol Chem 264, 1172511731.CrossRefGoogle ScholarPubMed
89Albrieux, M, Lee, HC & Villaz, M (1998) Calcium signaling by cyclic ADP-ribose, NAADP, and inositol trisphosphate are involved in distinct functions in ascidian oocytes. J Biol Chem 273, 1456614574.CrossRefGoogle ScholarPubMed
90Moccia, F, Nusco, GA, Lim, D, Ercolano, E, Gragnaniello, G, Brown, ER & Santella, L (2003) Ca2+ signalling and membrane current activated by cADPr in starfish oocytes. Pflugers Arch 446, 541552.Google Scholar
91Khoo, KM, Han, MK, Park, JB, Chae, SW, Kim, UH, Lee, HC, Bay, BH & Chang, CF (2000) Localization of the cyclic ADP-ribose-dependent calcium signaling pathway in hepatocyte nucleus. J Biol Chem 275, 2480724817.Google Scholar
92Yusufi, AN, Cheng, J, Thompson, MA, Dousa, TP, Warner, GM, Walker, HJ & Grande, JP (2001) cADP-ribose/ryanodine channel/Ca2+-release signal transduction pathway in mesangial cells. Am J Physiol Renal Physiol 281, F91F102.Google Scholar
93Guse, AH (2004) Biochemistry, biology, and pharmacology of cyclic adenosine diphosphoribose (cADPR). Curr Med Chem 11, 847855.CrossRefGoogle ScholarPubMed
94Lee, HC, Aarhus, R, Graeff, R, Gurnack, ME & Walseth, TF (1994) Cyclic ADP ribose activation of the ryanodine receptor is mediated by calmodulin. Nature 370, 307309.Google Scholar
95Guse, AH, Tsygankov, AY, Weber, K & Mayr, GW (2001) Transient tyrosine phosphorylation of human ryanodine receptor upon T cell stimulation. J Biol Chem 276, 3472234727.Google Scholar
96Beutner, G, Sharma, VK, Giovannucci, DR, Yule, DI & Sheu, SS (2001) Identification of a ryanodine receptor in rat heart mitochondria. J Biol Chem 276, 2148221488.Google Scholar
97Khoo, KM & Chang, CF (2002) Identification and characterization of nuclear CD38 in the rat spleen. Int J Biochem Cell Biol 34, 4354.Google Scholar
98Adebanjo, OA, Anandatheerthavarada, HK, Koval, AP, et al. . (1999) A new function for CD38/ADP-ribosyl cyclase in nuclear Ca2+ homeostasis. Nat Cell Biol 1, 409414.Google Scholar
99Schwarzmann, N, Kunerth, S, Weber, K, Mayr, GW & Guse, AH (2002) Knock-down of the type 3 ryanodine receptor impairs sustained Ca2+ signaling via the T cell receptor/CD3 complex. J Biol Chem 277, 5063650642.Google Scholar
100Lukyanenko, V, Gyorke, I, Wiesner, TF & Gyorke, S (2001) Potentiation of Ca(2+) release by cADP-ribose in the heart is mediated by enhanced SR Ca(2+) uptake into the sarcoplasmic reticulum. Circ Res 89, 614622.Google Scholar
101Itoh, M, Ishihara, K, Tomizawa, H, Tanaka, H, Kobune, Y, Ishikawa, J, Kaisho, T & Hirano, T (1994) Molecular cloning of murine BST-1 having homology with CD38 and Aplysia ADP-ribosyl cyclase. Biochem Biophys Res Commun 203, 13091317.Google Scholar
102Kim, H, Jacobson, EL & Jacobson, MK (1993) Synthesis and degradation of cyclic ADP-ribose by NAD glycohydrolases. Science 261, 13301333.CrossRefGoogle ScholarPubMed
103Ceni, C, Pochon, N, Villaz, M, Muller-Steffner, H, Schuber, F, Baratier, J, De Waard, M, Ronjat, M & Moutin, MJ (2006) The CD38-independent ADP-ribosyl cyclase from mouse brain synaptosomes: a comparative study of neonate and adult brain. Biochem J 395, 417426.Google Scholar
104De Flora, A, Guida, L, Franco, L & Zocchi, E (1997) The CD38/cyclic ADP-ribose system: a topological paradox. Int J Biochem Cell Biol 29, 11491166.CrossRefGoogle ScholarPubMed
105Bruzzone, S, Guida, L, Zocchi, E, Franco, L & De Flora, A (2001) Connexin 43 hemi channels mediate Ca2+-regulated transmembrane NAD+ fluxes in intact cells. FASEB J 15, 1012.Google Scholar
106Franco, L, Guida, L, Bruzzone, S, Zocchi, E, Usai, C & De Flora, A (1998) The transmembrane glycoprotein CD38 is a catalytically active transporter responsible for generation and influx of the second messenger cyclic ADP-ribose across membranes. FASEB J 12, 15071520.Google Scholar
107De Flora, A, Zocchi, E, Guida, L, Franco, L & Bruzzone, S (2004) Autocrine and paracrine calcium signaling by the CD38/NAD+/cyclic ADP-ribose system. Ann N Y Acad Sci 1028, 176191.Google Scholar
108Guida, L, Bruzzone, S, Sturla, L, Franco, L, Zocchi, E & De Flora, A (2002) Equilibrative and concentrative nucleoside transporters mediate influx of extracellular cyclic ADP-ribose into 3T3 murine fibroblasts. J Biol Chem 277, 4709747105.Google Scholar
109Podestà, M, Zocchi, E, Pitto, A, et al. . (2000) Extracellular cyclic ADP-ribose increases intracellular free calcium concentration and stimulates proliferation of human hemopoietic progenitors. FASEB J 14, 680690.Google Scholar
110Franco, L, Zocchi, E, Usai, C, Guida, L, Bruzzone, S, Costa, A & De Flora, A (2001) Paracrine roles of NAD+ and cyclic ADP-ribose in increasing intracellular calcium and enhancing cell proliferation of 3T3 fibroblasts. J Biol Chem 276, 2164221648.Google Scholar
111Verderio, C, Bruzzone, S, Zocchi, E, Fedele, E, Schenk, U, De Flora, A & Matteoli, M (2001) Evidence of a role for cyclic ADP-ribose in calcium signalling and neurotransmitter release in cultured astrocytes. J Neurochem 78, 646657.Google Scholar
112Franco, L, Bruzzone, S, Song, P, Guida, L, Zocchi, E, Walseth, TF, Crimi, E, Usai, C, De Flora, A & Brusasco, V (2001) Extracellular cyclic ADP-ribose potentiates ACh-induced contraction in bovine tracheal smooth muscle. Am J Physiol Lung Cell Mol Physiol 280, L98L106.Google Scholar
113Sun, L, Adebanjo, OA, Koval, A, et al. . (2002) A novel mechanism for coupling cellular intermediary metabolism to cytosolic Ca2+ signaling via CD38/ADP-ribosyl cyclase, a putative intracellular NAD+ sensor. FASEB J 16, 302314.Google Scholar
114Munshi, CB, Graeff, R & Lee, HC (2002) Evidence for a causal role of CD38 expression in granulocytic differentiation of human HL-60 cells. J Biol Chem 277, 4945349458.Google Scholar
115Ceni, C, Pochon, N, Brun, V, et al. . (2003) CD38-dependent ADP-ribosyl cyclase activity in developing and adult mouse brain. Biochem J 370, 175183.Google Scholar
116Yamada, M, Mizuguchi, M, Otsuka, N, Ikeda, K & Takahashi, H (1997) Ultrastructural localization of CD38 immunoreactivity in rat brain. Brain Res 756, 5260.Google Scholar
117Mizuguchi, M, Otsuka, N, Sato, M, Ishii, Y, Kon, S, Yamada, M, Nishina, H, Katada, T & Ikeda, K (1995) Neuronal localization of CD38 antigen in the human brain. Brain Res 697, 235240.Google Scholar
118Malenka, RC & Bear, MF (2004) LTP and LTD: an embarrassment of riches. Neuron 44, 521.Google Scholar
119Malinow, R, Schulman, H & Tsien, RW (1989) Inhibition of postsynaptic PKC or CaMKII blocks induction but not expression of LTP. Science 245, 862866.Google Scholar
120Silva, AJ, Wang, Y, Paylor, R, Wehner, JM, Stevens, CF & Tonegawa, S (1992) α Calcium/calmodulin kinase II mutant mice: deficient long-term potentiation and impaired spatial learning. Cold Spring Harb Symp Quant Biol 57, 527539.Google Scholar
121Mulkey, RM, Endo, S, Shenolikar, S & Malenka, RC (1994) Involvement of a calcineurin/inhibitor-1 phosphatase cascade in hippocampal long-term depression. Nature 369, 486488.Google Scholar
122Mulkey, RM, Herron, CE & Malenka, RC (1993) An essential role for protein phosphatases in hippocampal long-term depression. Science 261, 10511055.Google Scholar
123Sabatini, BL, Oertner, TG & Svoboda, K (2002) The life cycle of Ca(2+) ions in dendritic spines. Neuron 33, 439452.CrossRefGoogle ScholarPubMed
124Nicoll, RA, Oliet, SH & Malenka, RC (1998) NMDA receptor-dependent and metabotropic glutamate receptor-dependent forms of long-term depression coexist in CA1 hippocampal pyramidal cells. Neurobiol Learn Mem 70, 6272.Google Scholar
125O'Mara, SM, Rowan, MJ & Anwyl, R (1995) Dantrolene inhibits long-term depression and depotentiation of synaptic transmission in the rat dentate gyrus. Neuroscience 68, 621624.CrossRefGoogle ScholarPubMed
126Wang, Y, Rowan, MJ & Anwyl, R (1997) Induction of LTD in the dentate gyrus in vitro is NMDA receptor independent, but dependent on Ca2+ influx via low-voltage-activated Ca2+ channels and release of Ca2+ from intracellular stores. J Neurophysiol 77, 812825.Google Scholar
127Reyes, M & Stanton, PK (1996) Induction of hippocampal long-term depression requires release of Ca2+ from separate presynaptic and postsynaptic intracellular stores. J Neurosci 16, 59515960.Google Scholar
128Etherington, SJ & Everett, AW (2004) Postsynaptic production of nitric oxide implicated in long-term depression at the mature amphibian (Bufo marinus) neuromuscular junction. J Physiol 559, 507517.Google Scholar
129Lu, YF, Kandel, ER & Hawkins, RD (1999) Nitric oxide signaling contributes to late-phase LTP and CREB phosphorylation in the hippocampus. J Neurosci 19, 1025010261.Google Scholar
130Raymond, CR & Redman, SJ (2002) Different calcium sources are narrowly tuned to the induction of different forms of LTP. J Neurophysiol 88, 249255.Google Scholar
131Shimuta, M, Yoshikawa, M, Fukaya, M, Watanabe, M, Takeshima, H & Manabe, T (2001) Postsynaptic modulation of AMPA receptor-mediated synaptic responses and LTP by the type 3 ryanodine receptor. Mol Cell Neurosci 17, 921930.Google Scholar
132Futatsugi, A, Kato, K, Ogura, H, Li, ST, Nagata, E, Kuwajima, G, Tanaka, K, Itohara, S & Mikoshiba, K (1999) Facilitation of NMDAR-independent LTP and spatial learning in mutant mice lacking ryanodine receptor type 3. Neuron 24, 701713.Google Scholar
133Balschun, D, Wolfer, DP, Bertocchini, F, Barone, V, Conti, A, Zuschratter, W, Missiaen, L, Lipp, HP, Frey, JU & Sorrentino, V (1999) Deletion of the ryanodine receptor type 3 (RyR3) impairs forms of synaptic plasticity and spatial learning. EMBO J 18, 52645273.CrossRefGoogle ScholarPubMed
134Morris, RG, Garrud, P, Rawlins, JN & O'Keefe, J (1982) Place navigation impaired in rats with hippocampal lesions. Nature 297, 681683.Google Scholar
135Morris, RGM (1981) Spatial localization does not require the presence of local cues. Learn Mem 12, 239260.Google Scholar
136Zhao, W, Meiri, N, Xu, H, Cavallaro, S, Quattrone, A, Zhang, L & Alkon, DL (2000) Spatial learning induced changes in expression of the ryanodine type II receptor in the rat hippocampus. FASEB J 14, 290300.Google Scholar
137Young, GS, Jacobson, EL & Kirkland, JB (2007) Water maze performance in young male Long–Evans rats is inversely affected by dietary intakes of niacin and may be linked to levels of the NAD+ metabolite cADPR. J Nutr 137, 10501057.Google Scholar
138Warren, SG & Juraska, JM (1997) Spatial and nonspatial learning across the rat estrous cycle. Behav Neurosci 111, 259266.Google Scholar
139Shibata, K & Kondo, T (1993) Effects of progesterone and estrone on the conversion of tryptophan to nicotinamide in rats. Biosci Biotechnol Biochem 57, 18901893.Google Scholar
140D'Hooge, R & De Deyn, PP (2001) Applications of the Morris water maze in the study of learning and memory. Brain Res Brain Res Rev 36, 6090.Google Scholar
141Akiyama, K, Ichinose, S, Omori, A, Sakurai, Y & Asou, H (2002) Study of expression of myelin basic proteins (MBPs) in developing rat brain using a novel antibody reacting with four major isoforms of MBP. J Neurosci Res 68, 1928.Google Scholar
142Reeves, PG, Nielsen, FH & Fahey, GC Jr (1993) AIN-93 purified diets for laboratory rodents: final report of the American Institute of Nutrition ad hoc writing committee on the reformulation of the AIN-76A rodent diet. J Nutr 123, 19391951.Google Scholar
143Spector, R (1979) Niacin and niacinamide transport in the central nervous system. in vivo studies. J Neurochem 33, 895904.Google Scholar
144Pozzilli, P, Browne, PD & Kolb, H (1996) Meta-analysis of nicotinamide treatment in patients with recent-onset IDDM. The Nicotinamide Trialists. Diabetes Care 19, 13571363.CrossRefGoogle ScholarPubMed
145Kirkland, JB & Rawling, JM (2001) Niacin. In Handbook of Vitamins, 3rd ed., pp. 213254 [Rucker, RB, Zempleni, J, Suttie, JW and McCormick, DB, editors]. New York: Marcel Dekker, Inc.Google Scholar
146Ahmed, MG, Bedi, KS, Warren, MA & Kamel, MM (1987) Effects of a lengthy period of undernutrition from birth and subsequent nutritional rehabilitation on the synapse: granule cell neuron ratio in the rat dentate gyrus. J Comp Neurol 263, 146158.Google Scholar
147Jacobson, EL, Dame, AJ, Pyrek, JS & Jacobson, MK (1995) Evaluating the role of niacin in human carcinogenesis. Biochimie 77, 394398.Google Scholar
148Young, GS, Choleris, E, Lund, FE & Kirkland, JB (2006) Decreased cADPR and increased NAD(+) in the Cd38( − / − ) mouse. Biochem Biophys Res Commun 346, 188192.CrossRefGoogle Scholar
149Young, GS, Choleris, E, Lund, F & Kirkland, JB (2008) Like niacin deficient rats, Cd38− / −  mice show improved performance in the water maze. Curr Topics in Nutr Res (In the Press).Google Scholar
150Deshpande, DA, White, TA, Guedes, AG, Milla, C, Walseth, TF, Lund, FE & Kannan, MS (2005) Altered airway responsiveness in CD38-deficient mice. Am J Respir Cell Mol Biol 32, 149156.CrossRefGoogle ScholarPubMed
151Kato, I, Yamamoto, Y, Fujimura, M, Noguchi, N, Takasawa, S & Okamoto, H (1999) CD38 disruption impairs glucose-induced increases in cyclic ADP-ribose, [Ca2+]i, and insulin secretion. J Biol Chem 274, 18691872.Google Scholar
152Sun, L, Iqbal, J, Dolgilevich, S, et al. . (2003) Disordered osteoclast formation and function in a CD38 (ADP-ribosyl cyclase)-deficient mouse establishes an essential role for CD38 in bone resorption. FASEB J 17, 369375.Google Scholar
153Partida-Sanchez, S, Randall, TD & Lund, FE (2003) Innate immunity is regulated by CD38, an ecto-enzyme with ADP-ribosyl cyclase activity. Microbes Infect 5, 4958.Google Scholar
154Cockayne, DA, Muchamuel, T, Grimaldi, JC, Muller-Steffner, H, Randall, TD, Lund, FE, Murray, R, Schuber, F & Howard, MC (1998) Mice deficient for the ecto-nicotinamide adenine dinucleotide glycohydrolase CD38 exhibit altered humoral immune responses. Blood 92, 13241333.Google Scholar
155Anonymous (1997) Mutant mice and neuroscience: recommendations concerning genetic background. Banbury Conference on genetic background in mice. Neuron 19, 755759.Google Scholar
156Crusio, WE (2004) Flanking gene and genetic background problems in genetically manipulated mice. Biol Psychiatry 56, 381385.Google Scholar
157Gerlai, R (1996) Gene-targeting studies of mammalian behavior: is it the mutation or the background genotype? Trends Neurosci 19, 177181.Google Scholar
158Crawley, JN, Belknap, JK, Collins, A, et al. . (1997) Behavioral phenotypes of inbred mouse strains: implications and recommendations for molecular studies. Psychopharmacology (Berl) 132, 107124.Google Scholar
159Young, GS & Kirkland, JB (2006) Modifications to increase the efficiency of the fluorimetric cycling assay for cyclic ADP-ribose. Comb Chem High Throughput Screen 9, 633637.Google Scholar
160Schuber, F & Lund, FE (2004) Structure and enzymology of ADP-ribosyl cyclases: conserved enzymes that produce multiple calcium mobilizing metabolites. Curr Mol Med 4, 249261.Google Scholar
161Gallagher, M, Burwell, R & Burchinal, M (1993) Severity of spatial learning impairment in aging: development of a learning index for performance in the Morris water maze. Behav Neurosci 107, 618626.Google Scholar
162Takeshima, H, Ikemoto, T, Nishi, M, et al. . (1996) Generation and characterization of mutant mice lacking ryanodine receptor type 3. J Biol Chem 271, 1964919652.Google Scholar
163Higashida, H, Hashii, M, Yokoyama, S, Hoshi, N, Chen, XL, Egorova, A, Noda, M & Zhang, JS (2001) Cyclic ADP-ribose as a second messenger revisited from a new aspect of signal transduction from receptors to ADP-ribosyl cyclase. Pharmacol Ther 90, 283296.Google Scholar
164Yang, J, Han, H, Cao, J, Li, L & Xu, L (2006) Prenatal stress modifies hippocampal synaptic plasticity and spatial learning in young rat offspring. Hippocampus 16, 431436.Google Scholar
165Xiong, W, Yang, Y, Cao, J, Wei, H, Liang, C, Yang, S & Xu, L (2003) The stress experience dependent long-term depression disassociated with stress effect on spatial memory task. Neurosci Res 46, 415421.Google Scholar
Figure 0

Fig. 1 Chemical structures of niacin compounds: (a) nicotinamide; (b) nicotinic acid; (c) nicotinamide adenine dinucleotide (NAD+); (d) nicotinamide adenine dinucleotide phosphate (NADP+).

Figure 1

Fig. 2 Structure and origin of cyclic adenosine diphosphate ribose.

Figure 2

Table 1 Characteristics of inositol 1,4,5-triphosphate (IP3), cyclic adenosine diphosphate ribose (cADPR) and nicotinic acid adenine dinucleotide phosphate (NAADP)

Figure 3

Table 2 Intracellular effects of cyclic adenosine diphosphate ribose

Figure 4

Table 3 Composition of experimental diets (g/kg diet)

Figure 5

Fig. 3 (a) Cumulative error of niacin-deficient (–●–) and pair-fed (–○–) rats in the water maze. Rats were tested in three daily trials across 6 d with an inter-trial interval of 2 h. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means (n 8), with their standard errors represented by vertical bars. * Mean value was significantly different from that of the pair-fed rats (P ≤ 0·05). (b) Cumulative error of niacin-deficient (–●–; n 9) and partially feed-restricted (–○–; n 8) rats in the water maze. Rats were tested in three daily trials across 6 d with an inter-trial interval of 2 h. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means, with their standard errors represented by vertical bars. * Mean value was significantly different from that of the partially feed-restricted rats (P ≤ 0·05). (c) Cumulative error of niacin-deficient (–●–) and niacin-recovered (–○–) rats during reversal training in the water maze. Rats were tested in three daily trials across 4 d with an inter-trial interval of 2 h. The reversal training followed an initial acquisition phase in the water maze and 4 d of niacin refeeding. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means (n 9), with their standard errors represented by vertical bars. * Mean value was significantly different from that of the niacin-recovered rats (P ≤ 0·05). (d) Cumulative error of niacin-supplemented (–●–; n 18) and control (–○–; n 15) rats in the water maze. Rats were tested in three daily trials across 6 d with an inter-trial interval of 2 h. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means, with their standard errors represented by vertical bars. * Mean value was significantly different from that of the control rats (P ≤ 0·05). (e) Proximity averages to the platform during hidden platform testing by Cd38− / −  (–●–) and wild-type (–○–) mice across 7 d of testing. Mice were tested in three daily trials across 6 d with an inter-trial interval of 2 h. The results of the three daily trials were averaged to give a mean value for each day of testing. Values are means (n 10), with their standard errors represented by vertical bars. * Mean value was significantly different from that of the wild-type rats (P ≤ 0·05). Fig. 3(a–d) were originally published in Young et al. (2007)(137). Fig. 3(e) was originally published in Young et al. (2008)(149).

Figure 6

Table 4 Brain NAD+ and cyclic adenosine diphosphate ribose (cADPR) in rats with differing niacin intakes and in Cd38−/− mice (nmol/g tissue) (Mean values with their standard errors)