Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-25T00:07:36.189Z Has data issue: false hasContentIssue false

Modulating adult neurogenesis through dietary interventions

Published online by Cambridge University Press:  01 July 2016

Christine Heberden*
Affiliation:
Micalis Institute, INRA, AgroParisTech, Université Paris-Saclay, 78350 Jouy-en-Josas, France
*
Corresponding author: Christine Heberden, email Christine.heberden@jouy.inra.fr
Rights & Permissions [Opens in a new window]

Abstract

Three areas in the brain continuously generate new neurons throughout life: the subventricular zone lining the lateral ventricles, the dentate gyrus in the hippocampus and the median eminence in the hypothalamus. These areas harbour neural stem cells, which contribute to neural repair by generating daughter cells that then become functional neurons or glia. Impaired neurogenesis leads to detrimental consequences, such as depression, decline of cognitive abilities and obesity. Adult neurogenesis is a versatile process that can be modulated either positively or negatively by many effectors, external or endogenous. Diet can modify neurogenesis both ways, either directly by ways of food-borne molecules, or possibly by the modifications induced on gut microbiota composition. It is therefore critical to define dietary strategies optimal for the maintenance of the stem cell pools.

Type
Review Article
Copyright
Copyright © The Author 2016 

Introduction

Like any other organ, the brain possesses regenerative capacities, and the process of adult neurogenesis, since its existence was proven, has been the subject of intense investigation. The discovery of neural stem cells (NSC) has also instigated new hopes for neuronal repair in regenerative medicine, and inspired studies designed to better understand the events and regulations involved in neurogenesis. Even if the understanding of this complex process is still uncompleted, the knowledge of the factors influencing neuron generation is more and more precise and detailed, and the ensemble of neurogenesis-related studies shows that neurogenesis is a highly malleable process, in which diet can play an important part.

Neurogenesis is restricted to specific areas

Compared with high-rate turnover tissues such as the intestine for example, the brain presents limited regenerative capacities, in terms of intensity as well as in terms of location. The stem cells in the brain comply with the requirements defining stem cells: they self-renew, and are multipotent, i.e. they can differentiate into functional neurons, astrocytes and oligodendrocytes. The NSC are localised in three areas: the subventricular zone (SVZ) lining the lateral ventricle, the subgranular zone of the dentate gyrus (DG) in the hippocampus and a recently acknowledged area of the hypothalamus, the median eminence (ME). In the parenchyma, some scattered dividing cells have also been described, but these dividing cells are progenitors and not stem cells, in the sense that they do not self-renew, and that their in vivo fate is limited to the glial lineage. They will not be considered here.

The NSC of the SVZ give rise to neuroblasts migrating to the olfactory bulb, where they differentiate into GABAergic olfactory bulb interneurons and integrate the existing neuronal networks( Reference Lazarini and Lledo 1 ). They take part in the maintenance of the network, the incorporation of new olfactory stimuli, or the olfactory memory. The fate of the NSC has been precisely documented in the SVZ: NSC (type B cells) derived from the embryonic radial glial cells divide slowly and give rise to transient-amplifying cells (type C cells), precursors of the neuroblasts (type A cells) which migrate into the rostral migratory stream to reach the olfactory bulb( Reference Lazarini and Lledo 1 Reference Doetsch 3 ). NSC have also the possibility to differentiate into astrocytes and oligodendrocytes( Reference Jessberger, Toni and Clemenson 4 , Reference Menn, Garcia-Verdugo and Yaschine 5 ).

In the DG of the hippocampus, the NSC are designated by different names, but the successive events are very similar: the NSC are named radial and horizontal type 1 NSC. They divide slowly into type 2 neuronal precursors, which cycle more intensely. The precursors differentiate into neuroblasts and are finally integrated into the network of the granular zone after 4 weeks of maturation as glutamatergic granule cells. The difference in this structure resides in the fact that the neuroblasts migrate shorter distances, and differentiate in the layer of subgranular cells. They establish connections with the pyramidal layer of cornu Ammonis (CA) 3 zone of the hippocampus( Reference van Praag, Schinder and Christie 6 , Reference Toni, Laplagne and Zhao 7 ). Correlative studies by ablation and deletion have established that the new neurons are involved in the maintenance of spatial memory, learning capacities, and the retention of new memories( Reference Zhao, Deng and Gage 8 ). There seems to be a correlation between the survival of newborn cells and spatial-memory performances( Reference Borcel, Perez-Alvarez and Herrero 9 , Reference Trouche, Bontempi and Roullet 10 ). DG neurogenesis has also been linked with anxiety and depression or depression susceptibility in humans. This aspect has been under intense focus after the groundbreaking studies showing that the anti-depressant fluoxetine could increase cell division in the DG, and that this effect could correspond to the time needed for the treatment to have an impact in the patients( Reference Encinas, Vaahtokari and Enikolopov 11 , Reference Lee, Kim and Yim 12 ).

If the existence of neurogenesis is now firmly established in rodents’ brains, the relevance of these observations in humans has been debated. In humans, SVZ neurogenesis is very active until 2 years of age, with large numbers of immature neuroblasts migrating in the rostral migratory stream, but is nearly extinct in adulthood( Reference Sanai, Nguyen and Ihrie 13 ). A similar observation has been made in non-human primates( Reference Wang, Liu and Liu 14 ). By contrast, human DG neurogenesis remains very active in adults: it has been estimated that 700 neurons/d are added in the DG( Reference Spalding, Bergmann and Alkass 15 ).

Another very important role of neurogenesis in both areas is that the generation of new neurons is involved in neuronal repair after ischaemic stroke. Neuroblasts from the SVZ migrate to the damaged areas such as the cortex and the striatum and serve as a replacement of the dead neurons( Reference Jin, Sun and Xie 16 , Reference Liu, Solway and Messing 17 ) and ischaemic stroke enhances also neurogenesis in the DG( Reference Tan, Preston and Wojtowicz 18 ). NSC transplants can survive and migrate into the ischaemic area( Reference Kelly, Bliss and Shah 19 ) and neuroblast transplantation helps the restoration of behaviour and motor skills in rodents( Reference Chintawar, Hourez and Ravella 20 Reference Nori, Okada and Yasuda 22 ).

There is a great similarity in the markers of the NSC and progenitors from both areas: nestin and Sox2 are the classical markers for the NSC, and are not expressed anymore along the line of differentiation, polysialylated neuronal cell adhesion molecule (PSA-NCAM) and doublecortin are the most commonly used markers for neuroblasts, which once fully differentiated into neurons, can be labelled with NeuN or HuC/D. Using these markers enables to trace the fate of the newly generated cells in vivo.

The hypothalamus has been less studied. There have been reports showing the existence of dividing cells for long( Reference Xu, Tamamaki and Noda 23 ); but this structure has recently gained further interest because studies have pointed out that the neurogenesis in this area is modulated by high-fat diets, a point that will be discussed below. Tanycytes are glial cells that line the third ventricle. Four different types have been characterised, α 1 and 2, and β 1 and 2, the function and properties of which differ with their location along the ventricle( Reference Rodriguez, Blazquez and Pastor 24 ). The β2 tanycytes, at the bottom of the third ventricle in the ME region and the α2 subtype, bordering the lateral walls of the third ventricle in the paraventricular and arcuate nuclei, are now considered as NSC, because they proliferate and display markers such as nestin and Sox2( Reference Lee and Blackshaw 25 ). They migrate under intermediate differentiation into the arcuate nucleus (ARC) of the hypothalamus, and differentiate fully into pro-opiomelanocortin (POMC) neurons. POMC neurons are anorexigenic neurons; they decrease food intake through αMSH release, a proteolytic product of POMC. A fraction of these new neurons differentiate in a second wave into NPY and Agouti-related peptide (AgRP) neurons, which are orexigenic, and increase food intake( Reference McNay, Briancon and Kokoeva 26 , Reference Padilla, Carmody and Zeltser 27 ). The origin of the feeding circuitry lies in the embryonic period, and the remodelling is very intense postnatally, and up until adolescence in the rodent( Reference Tan, Preston and Wojtowicz 18 ) but persists into adulthood( Reference Li, Tang and Cai 28 ). Disruption of the neurogenic process leads to modifications of food intake: an acute degeneration of AgRP neurons leads to a severe anorexia in mice, while a slow and progressive deletion does not have a significant effect on food intake, because neurogenesis can compensate and replace the degenerated neurons( Reference Pierce and Xu 29 ). Therefore, the loss of POMC neurons in ARC could favour the development of obesity.

The neurogeneses in the three areas present similarities on several aspects: (1) the stem cells display markers specific to the glial lineage( Reference Barry, Pakan and McDermott 30 Reference Pinto and Gotz 32 ); (2) the succession of events involved, i.e. proliferation followed by migration and differentiation; and (3) the nature of the markers. Yet, because each area is also a particular niche, their regulations can present some differences.

Regulation and effectors

As mentioned above, NSC are not scattered in the parenchyma, but are concentrated in particular areas named ‘niches’ which provides the environment necessary to protect and preserve the cell pool. The niches are localised in proximity to ventricles (SVZ and ME) and/or capillaries (DG). Therefore, they are in close contact with effectors and the NSC pool can react to signalling modifications, either by dividing or staying quiescent. The signals can be endogenous; they include hormones, growth factors or neurotrophins. They can also be external, and that is how diet and nutrition can influence the process.

Neurotrophins and growth factors are positive regulators of neurogenesis: brain-derived neurotrophic factor (BDNF), insulin-like growth factor or vascular endothelial growth factor all increase cell proliferation, and/or promote cell survival or cell differentiation in the DG( Reference Bergami, Rimondini and Santi 33 , Reference Sairanen, Lucas and Ernfors 34 ). A broad range of hormones can also act on neurogenesis: oestrogen, progesterone and dehydroepiandrosterone promote DG neurogenesis( Reference Li and Shen 35 , Reference Sicard, Ehrhart-Bornstein and Corbeil 36 ), while corticosteroids, the last messengers of the hypothalamic–pituitary–adrenal axis, diminish cell proliferation and differentiation( Reference Denis, Potier and Vancassel 37 ). Peptide hormones can also modify neurogenesis: leptin promotes cell proliferation in the DG( Reference Garza, Guo and Zhang 38 ) and insulin promotes neuronal differentiation( Reference Han, Wang and Xiao 39 ). Gut peptides can modify cell renewal: ghrelin, peptide YY and cholecystokinin can increase cell proliferation in the DG and ghrelin acts positively on all steps of neurogenesis in the SVZ( Reference Li, Kim and Kim 40 , Reference Stanic, Paratcha and Ledda 41 ). Chronic injections of glucagon-like peptide-1 increased neurogenesis in a model of Alzheimer’s disease( Reference Parthsarathy and Holscher 42 ); the effect of satiety peptides have been described on DG neurogenesis so far, and their effect on the proliferation and differentiation of β2 tanycytes is not known. It would be interesting to examine this point. Indeed, although the mechanisms involved in the three areas unroll very similarly, they are not necessarily modified by the same factors: significant differences in the length and importance of the successive steps may be responsible for the differences in sensitivity towards effectors. For instance, prolactin enhances SVZ neurogenesis and thereby increases the olfactory potential of the mother in preparation for maternal behaviour but DG neurogenesis is not modified by prolactin( Reference Larsen and Grattan 43 ). Conversely, BDNF increases cell proliferation in the DG, but does not affect cell proliferation in the SVZ( Reference Galvao, Garcia-Verdugo and Alvarez-Buylla 44 ).

A common feature of neurogenesis in the three areas is the vulnerability to the ageing process. Following embryonic neurogenesis, proliferation is very active in the postnatal period and decreases significantly until mid-age, then keeps decreasing at a slower pace during senescence in the hippocampus( Reference Bondolfi, Ermini and Long 45 , Reference Kuhn, Dickinson-Anson and Gage 46 ). Recently, a study by Zhang et al. ( Reference Zhang, Li and Purkayastha 47 ) showed that inflammation in the hypothalamus increases with age, as measured by NF-κB activation in the microglia; this ageing process negatively regulated gonadotropin-releasing hormone (GnRH) synthesis, and GnRH delivery into the third ventricle could reverse age-induced neurogenesis decline.

Dietary influences

Environmental factors are efficient actors in the regulation of neurogenesis and diet or dietary intakes take an important part in these interactions. Several aspects can be considered: the quantity of ingested food or its composition, as developed below, and even its texture: rat hippocampus proliferation was decreased in animals fed soft food, and after being fed a powdered diet for 1 year, mice had lower counts of pyramidal neurons in the CA1 and CA3 regions of the hippocampus than mice fed a solid diet( Reference Aoki, Kimoto and Hori 48 , Reference Nose-Ishibashi, Watahiki and Yamada 49 ).

Energy restriction v. overnutrition

Dietary restriction achieved by a reduction of 20–40 % of energy intake is reported to promote many health benefits, and to extend lifespan in many organisms, including mammals. The mechanism involved is thought to be linked to a reduction in oxidative stress( Reference Canto and Auwerx 50 Reference Ruetenik and Barrientos 52 ). In the brain, dietary restriction increases the number of newborn neurons in the DG of male rats fed every other day. A similar observation has been made in mice fed according to the same schedule: proliferation in the DG was not modified, but the rate of the survival of the new neurons was increased. This has been put in relation with higher levels of BDNF in the CA1 of the hippocampus( Reference Lee, Duan and Long 53 , Reference Lee, Duan and Mattson 54 ).

Conversely, in male adult rats, overnutrition resulting from the ingestion of high-sugar( Reference Van der Borght, Köhnke and Göransson 55 ) or high-fat/high-energy diets leads to a reduction of NSC proliferation and neuron generation in the DG( Reference Lindqvist, Mohapel and Bouter 56 )and ARC( Reference Chintawar, Hourez and Ravella 20 ). However, a high-fat diet and obesity make an impact on cell proliferation in the brain selectively( Reference Rivera, Romero-Zerbo and Pavon 57 , Reference Yon, Mauger and Pickavance 58 ). In the ME, the consequences of a high-fat diet seem less consistent: a study showed a decrease in cell generation only after a long-term ingestion of a high-fat diet (4 months) in male adult rats. Another group, after a much more limited period of high-fat feeding (1 month), showed an opposite effect on cell proliferation, but this group studied young rats right after weaning( Reference Lee, Bedont and Pak 59 ). The 4-month feeding resulted in a diminution of cell proliferation, diminution of cell survival, with a decrease more important in the neuronal lineage, compared with the glial lineage; the effect on neurogenesis was connected with an increase in TNFα and IL-1β concentrations, resulting from an activation of the NF-κB pathway. This activation promoted a higher rate of apoptosis in the ARC, leading to a moderate loss of POMC neurons, while NPY neurons remained unchanged( Reference Chintawar, Hourez and Ravella 20 ). In young adult mice, a high-fat diet (60 % fat) resulted in a decrease in both sexes in ARC neurogenesis, but enhanced ME neurogenesis in female mice( Reference Lee, Yoo and Pak 60 ). A maternal diet rich in lipids and energy (32 % fat) also impairs the hippocampal and hypothalamic neurogenesis of young mouse offspring, with negative consequences on learning abilities and energy intake control( Reference Tozuka, Kumon and Wada 61 ). A concordant result has been established in patients, linking the Western diet and a smaller left hippocampus( Reference Jacka, Cherbuin and Anstey 62 ).

It is interesting to underscore that the results obtained after long-term high-fat feeding paralleled the consequences of the ageing effect: very similarly, ageing results in cytokine elevation and NF-κB activation in the hypothalamus( Reference Sicard, Ehrhart-Bornstein and Corbeil 36 ).

Consequently, food-borne molecules known for their anti-inflammatory and antioxidative properties could be considered valuable in view of dietary strategies aimed at restoring neurogenesis capacities.

Dietary interventions

The present paragraph will discuss the possibility of dietary strategies or supplementations using food-borne molecules with a focus on their role on neurogenesis, knowing that the effects of these molecules cover a more complex and much broader field than mere neuron generation.

Polyphenols are such compounds. They are known for their antioxidant and anti-inflammatory properties( Reference Rendeiro, Guerreiro and Williams 63 ). Flavonoids are a class of polyphenols which have attracted a considerable interest. Dietary supplementations of compounds such as oroxylin A, baicalin or heptamethoxyflavone, a citrus flavone, can increase cell proliferation in the hippocampus of mice following global cerebral ischaemia( Reference Lee, Kim and Lee 64 Reference Okuyama, Shimada and Kaji 66 ), and baicalein reduces the impairment provoked by γ-ray irradiation( Reference Oh, Park and Jang 67 ). They can also play a neuroprotective role, favouring neurite outgrowth and neural differentiation in vitro ( Reference Nakajima, Niisato and Marunaka 68 , Reference Oberbauer, Urmann and Steffenhagen 69 ). Resveratrol, a phenol present in wine, red grapes and groundnuts, can increase cell proliferation in prenatal stress or in mice with chronic fatigue syndrome( Reference Madhyastha, Sekhar and Rao 70 , Reference Moriya, Chen and Yamakawa 71 ), and the supplementation also led to a decrease in inflammation and improvement of cognitive functions. Low intakes of curcumin, a phenol present in curry powder, increased cell proliferation in the hippocampus( Reference Kim, Son and Park 72 ). Curcumin also has neuroprotective properties, as shown in aged rats: it can reverse the effects of age and ischaemia, and promote cell repair in spinal cord injury( Reference Dong, Zeng and Mitchell 73 , Reference Wang, Zu and Li 74 ).

n-3 PUFA have also generated much interest, being a major constituent of brain structures. Insufficient brain DHA, the main n-3 PUFA in cell membranes, has been associated with memory impairment, emotional disturbances and altered brain processes in experimental studies. Epidemiological studies have revealed that low n-3 PUFA is correlated with cognitive or behavioural defects during early development( Reference Coti Bertrand, O’Kusky and Innis 75 ), adulthood and ageing( Reference Latour, Grintal and Champeil-Potokar 76 ). Among many other roles, n-3 PUFA can improve hippocampal neurogenesis by enhancing cell proliferation and differentiation in adult rats( Reference Calderon and Kim 77 Reference He, Qu and Cui 79 ), and reverse the decline observed in aged rats under a short supplementation( Reference Dyall, Michael and Michael-Titus 80 ). In 19-month-old mice, 8 weeks of n-3 supplementation led to an increase in hippocampal neurogenesis and in arborisation of newborn neurons( Reference Cutuli, De Bartolo and Caporali 81 ). They also affect deeply the physiology of rat NSC, by enhancing their proliferation and differentiation, and modifying their trancriptome( Reference Goustard-Langelier, Koch and Lavialle 82 ).

Most of the studies described above refer to hippocampal neurogenesis, because until now, this aspect was the most relevant to the human situation. It would be now interesting to test these compounds on hypothalamic neurogenesis, particularly to check if they could reverse some of the aspects resulting from obesity, for instance.

Mechanisms involved

From the examples, a general outline can be drawn, showing that molecules or events that increase inflammation tend to lower neurogenesis, while anti-inflammatory molecules or pathways restore it. Therefore, the mediators of the inflammatory pathway represent the central targets and the balance upon which the different effectors can impinge. We will develop here some examples of these targets, such as the cAMP response element binding protein (CREB), BDNF, NF-κB, sirtuin 1 (Sirt1), and the energy-sensing kinase AMP-activated protein kinase (AMPK), and we will show how they all inter-relate.

Brain-derived neurotrophic factor and cAMP response element binding protein

BDNF is a well-recognised effector of neurogenesis. This neurotrophin is involved in brain development and neural plasticity. It mediates its effect through binding to the membrane receptor tropomyosin receptor kinase B (TrkB). The variations of BDNF in the hippocampus induced by environmental modifications and dietary interventions are directly correlated with the variations noted in neurogenesis: dietary restriction, which increases neurogenesis, and increases BDNF mRNA expression and protein( Reference Andrade, Mesquita and Assuncao 83 ). On the other hand, overnutrition decreases BDNF levels. It is here interesting to underline that BDNF, apart from its role on neurogenesis, acts also on food intake, since knock-out mice deficient in bdnf display hyperphagia and develop obesity( Reference Gray, Yeo and Cox 84 ). Very similarly, the effects of flavonoids on neurogenesis have been connected with parallel modifications of BDNF in whole brain or hippocampus: ingestion of naringin, oroxylin A or blueberry extracts leads to a significant increase in BDNF levels( Reference Jeon, Rhee and Seo 85 Reference Williams, El Mohsen and Vauzour 87 ). Resveratrol promotes BDNF synthesis from astroglia( Reference Zhang, Lu and Wu 88 ) and in the hippocampus( Reference Rahvar, Nikseresht and Shafiee 89 ), and by promoting BDNF synthesis reverses hippocampal atrophy in chronic fatigue mice and reverses the effect of mild stress on cognition( Reference Liu, Xie and Yang 90 , Reference Wang, Gu and Wang 91 ).

The effects of n-3 PUFA supplementation are similar: a large body of studies has unanimously shown that, in experimental models, the dietary supplementation of DHA or a mixture of n-3 PUFA could increase BDNF levels, or normalise them after mild brain injury in rats( Reference Wu, Ying and Gomez-Pinilla 92 ). DHA supplementation at 1·25 % could increase the effects of voluntary exercise in mice, with a parallel increase in BDNF in the brain. Quite interestingly, the benefits of n-3 PUFA were also transmitted by the maternal diet since prenatal n-3 PUFA supplementation to a micronutrient-imbalanced diet protected brain neurotrophins in both the cortex and hippocampus in the adult rat offspring( Reference Sable, Kale and Joshi 93 ).

In most cases the similarity of actions on BDNF synthesis could be related to the effects on synthesis of CREB and its phosphorylation. CREB is an ubiquitous transcription factor, highly expressed in the brain in which it is involved in many signalling pathways serving many functions, ranging from neuronal survival to memory formation( Reference Finkbeiner 94 ). Its activation through phosphorylation is the result of the action of the mitogen-activated protein kinases, or protein kinase C kinase. Once phosphorylated, CREB enters the nucleus where it binds to the CREB response element. One of its target gene is Bdnf. Experimental data showed that n-3 PUFA supplementation after traumatic brain injury enhanced CREB protein contents( Reference Wu, Ying and Gomez-Pinilla 95 ) and reduced oxidative damage in injured rats. Curcumin enhanced CREB expression in rat brain( Reference Kumar, Antony and Gireesh 96 , Reference Xu, Ku and Tie 97 ), and resveratrol increased the phosphorylated state of CREB after ischaemia( Reference Lin, Chen and Zhang 98 ).

Inflammatory NF-κB v. anti-inflammatory Sirt1 and AMP kinase

Another factor affecting neurogenesis is the balance between pro- and anti-inflammatory pathways.

Ageing, depression and obesity have all been related to low-grade neuroinflammation and its oxidative stress companion( Reference Tang, Purkayastha and Cai 99 , Reference Chaudhari, Talwar and Parimisetty 100 ). NF-κB is the major and central mediator of inflammation. This transcription factor, once activated, can induce the transcription of the inflammatory cytokines such as TNFα, monocyte chemoattractant protein-1 and IL-1β or IL-6( Reference Baker, Hayden and Ghosh 101 ). NF-κB activation has also been cited in the different situations leading to an increase or decrease of neurogenesis: all the signals, either endogenous such as age and stress, or external such as overnutrition, lead to an activation of NF-κB in the brain, either in the hippocampus or the hypothalamus, and decrease neurogenesis in these areas.

An endogenous inhibitor of NF-κB is Sirt1. Sirt1 is a deacetylase enzyme regulating energy metabolism and cell survival. Among other targets, the RelA/p65 component of the NF-κB complex is deacetylated by Sirt1 on Lys130, which inhibits the transcription ability of NF-κB( Reference Yeung, Hoberg and Ramsey 102 ). By this mechanism, Sirt1 can prevent the deleterious effects of a high-fat diet and protect from hepatic steatosis( Reference Pfluger, Herranz and Velasco-Miguel 103 ). Therefore, Sirt1 activators could also be valuable in preventing the inflammatory pathway. Dietary restriction is a strong inducer of Sirt1 expression and activity. The first compound detected among a screening of small molecules to be an inhibitor of Sirt1 was resveratrol( Reference Howitz, Bitterman and Cohen 104 ), and it was later demonstrated that resveratrol could mimic the effects of energy restriction and delay the ageing process through its up-regulation of Sirt1( Reference Wood, Rogina and Lavu 105 ). It is now well admitted that nutrients can modulate inflammation through Sirt1 involvement. Other flavonoids, such as quercetin or rutin for instance, can also up-regulate Sirt1 expression( Reference Davis, Murphy and Carmichael 106 , Reference Su, Yu and Chen 107 ). Because of its broad enzymic activity, Sirt1 can take part in many pathways. It has been demonstrated that Sirt1 activity could modify the differentiation fate of NSC, favouring astrogliosis to the expense of neurogenesis under oxidative stress( Reference Prozorovski, Schulze-Topphoff and Glumm 108 ). Sirt1 is also essential for cognition and neuronal plasticity( Reference Michán, Li and Chou 109 ). It is interesting to mention that the overexpression of Sirt1 stimulates BDNF expression( Reference Zocchi and Sassone-Corsi 110 ), and is neuroprotective in a model of Huntington’s disease( Reference Jiang, Wang and Fu 111 ). By its deacylating activity, Sirt1 is involved in a large range of effects. Among the numerous Sirt1 targets, we will focus on AMPK, because its involvement in NSC physiology and its regulation by dietary factors attract increasing interest.

AMPK is a fuel-sensing kinase that drives the cellular metabolism from catabolic to anabolic depending on the energy level of the cell. Its activity is up-regulated when the ATP levels of the cell are low. AMPK and Sirt1 share many cellular targets, are activated by the same stimuli, and also regulate each other( Reference Ruderman, Xu and Nelson 112 ). Experimental evidence from transgenic mice lacking the AMPKβ1 subunit demonstrates that this kinase is involved in NSC proliferation, since the mice display cerebral atrophy. AMPK could also improve neuronal survival under low energy conditions( Reference Dasgupta and Milbrandt 113 ). Several studies have shown that bioflavonoids can induce, in vitro and in vivo, an activation of the AMPK pathway. Naringenin, quercetin and quercetin 3-O-glycosides extracted from plants in muscle cells, or in the brain of old mice, up-regulate AMPK activity. Apigenin induces the activation in human keratinocytes and its activity is also up-regulated by resveratrol in mouse primary neurons( Reference Eid, Martineau and Saleem 114 Reference Vingtdeux, Giliberto and Zhao 118 ). Similarly, n-3 PUFA exert their protective and anti-inflammatory effects through the up-regulation of AMPK( Reference Wu, Ying and Gomez-Pinilla 95 ). The implication of Sirt1 has not been extensively investigated in these studies, but has been demonstrated in others( Reference Jing, Yao and Liu 119 , Reference Xue, Yang and Wang 120 ).

For clarity we have described the two pathways under two distinct paragraphs, but they actually are also intertwined and it is interesting here to notice that the mediators susceptible to act on neurogenesis are also found at the intersection of the fields of inflammation, energy homeostasis and neuronal plasticity, as recapitulated in Fig. 1.

Fig. 1 Adult neurogenesis is a flexible process, which can be modified by positive or negative influences. Pro-inflammatory molecules or mechanisms activate NF-κB, which induces the transcription of pro-inflammatory cytokines, while anti-inflammatory molecules by their action on sirtuin 1 (Sirt1) activate cAMP response element binding protein (CREB), brain-derived neurotrophic factor (BDNF) and AMP-activated protein kinase (AMPK) and restore cell renewal in the brain. The influence of gut microbiota is here hypothetical. For a colour figure, see the online version.

Rodent experimentation has amply demonstrated the disadvantages/benefits of diets/molecules on neurogenesis and their functional consequences such as the effects on learning, mood or energy homeostasis. The relevance of these studies to the human situation is hampered by the difficulty in measuring human neurogenesis so this point is mainly based on correlations: for instance, a recent study tested the supplementation of resveratrol in elderly and proved it efficient in protecting memory and glucose metabolism( Reference Witte, Kerti and Margulies 121 ). Similarly, components of the Mediterranean diet improved cognition in elderly( Reference Huhn, Kharabian Masouleh and Stumvoll 122 ). So data generated from animal experimentation have opened the way to many more studies on the usage of diets or natural compounds as therapeutic agents.

Gut microbiota: a possible effector?

Gut microbiota is now considered as a major factor in the modulation of the host’s physiology, including neural functions through the gut–brain axis. Seminal data obtained from the comparison of germ-free (GF) mice with their conventional counterparts have shown that the absence of intestinal microbiota was associated with a decreased anxiety but a stronger response to stress( Reference Diaz Heijtz, Wang and Anuar 123 ), but GF male F344 rats exhibited increased anxiety compared with SPF counterparts( Reference Crumeyrolle-Arias, Jaglin and Bruneau 124 ). The genetic background of the animals has been proposed as an important factor( Reference Crumeyrolle-Arias, Jaglin and Bruneau 124 ). GF mice display a lower abundance of BDNF in the hippocampus, and an increased monoamine turnover in the striatum( Reference Diaz Heijtz, Wang and Anuar 123 ). Another group has shown higher hippocampal concentrations of 5-hydroxytryptamine in male GF mice( Reference Clarke, Grenham and Scully 125 ).

A robust body of evidence has shown that gut microbiota composition differs between fat and lean animals, with a higher proportion of Firmicutes and a lower presence of Bacteroidetes in obese animals( Reference Tremaroli and Backhed 126 ). The modifications pointed out in animals are consistent with the modifications of the gut microbiota detected in obese patients( Reference Turnbaugh, Hamady and Yatsunenko 127 ). Moreover, microbiota alone seems able to modify the host’s metabolism, as exemplified by the recent clinical observation, resulting from a faecal transfer( Reference Alang and Kelly 128 ). The mechanisms involved are still under clarification, but hypotheses suggest that microbiota composition could either promote a more efficient energy harvest from the nutrients, or act through the lowering of the systemic inflammation level associated with obesity. Indeed, the inflammation develops with different kinetics in the organs, and the brain is the first one to be harmed, since a low level of inflammation is perceived after only 7 d of a high-fat diet( Reference Tan, Preston and Wojtowicz 18 ). Furthermore, the installation of obesity is accompanied with a deregulation of numerous hormones levels, such as insulin, leptin, adiponectin( Reference Song, Kang and Kim 129 ) and the gut satiety peptides such as glucagon-like peptide-1. We have seen in the paragraph above that these factors can positively modify neurogenesis, and, therefore, their deregulation could contribute to the development of a downward spiral aggravated by an impairment of neuronal repair. Recent data from obese and non-obese patients show that gut microbiota and brain microstructure are associated( Reference Fernandez-Real, Serino and Blasco 130 ). Whether ‘obese’ or ‘non-obese’ microbiota can modulate neuroinflammation or act directly through bacterial metabolites would be interesting questions to study, particularly in the hypothalamus, but also in the hippocampus, with regard to depression or decline of the cognitive abilities linked with obesity. Moreover, since the characteristics of the diet (for example, animal v. plant proteins) themselves can modify the composition of gut microbiota( Reference Zhu, Lin and Zhao 131 ), dietary recommendations should be made to drive the intestinal microbes’ mass to beneficial and optimal proportions.

Conclusion

Neurogenesis is an active component of brain physiology, and its malfunctioning can lead to adverse consequences such as depression, decline of learning abilities, and obesity. It is also a versatile and very sensitive process that can be modulated by many factors: if it can be impaired very easily, it can also be restored. Modifications in lifestyle can help keep the stem cell pools under optimal conditions, and nutrition is particularly helpful on this point. We have established here that most of the pernicious effects on neurogenesis are interrelated, and come from neuroinflammation. Keeping this inflammation down could help contribute to restore cell renewal; Fig. 1 sums up the detrimental effectors, and the possible solutions to counteract them. The experimental data have pointed out the favourable role of food components or molecules. A deeper knowledge of the possibilities linked with gut microbiota could also open new possibilities, and this could provide interesting alternatives to pharmaceutical treatments.

Acknowledgements

This research received no specific grant from any funding agency, commercial or not-for-profit sectors.

There are no conflicts of interest.

References

1. Lazarini, F & Lledo, PM (2011) Is adult neurogenesis essential for olfaction? Trends Neurosci 34, 2030.Google Scholar
2. Conover, JC & Allen, RL (2002) The subventricular zone: new molecular and cellular developments. Cell Mol Life Sci 59, 21282135.CrossRefGoogle ScholarPubMed
3. Doetsch, F (2003) A niche for adult neural stem cells. Curr Opin Genet Dev 13, 543550.Google Scholar
4. Jessberger, S, Toni, N, Clemenson, GD Jr, et al. (2008) Directed differentiation of hippocampal stem/progenitor cells in the adult brain. Nat Neurosci 11, 888893.CrossRefGoogle ScholarPubMed
5. Menn, B, Garcia-Verdugo, JM, Yaschine, C, et al. (2006) Origin of oligodendrocytes in the subventricular zone of the adult brain. J Neurosci 26, 79077918.Google Scholar
6. van Praag, H, Schinder, AF, Christie, BR, et al. (2002) Functional neurogenesis in the adult hippocampus. Nature 415, 10301034.Google Scholar
7. Toni, N, Laplagne, DA, Zhao, C, et al. (2008) Neurons born in the adult dentate gyrus form functional synapses with target cells. Nat Neurosci 11, 901907.Google Scholar
8. Zhao, C, Deng, W & Gage, FH (2008) Mechanisms and functional implications of adult neurogenesis. Cell 132, 645660.Google Scholar
9. Borcel, E, Perez-Alvarez, L, Herrero, AI, et al. (2008) Chronic stress in adulthood followed by intermittent stress impairs spatial memory and the survival of newborn hippocampal cells in aging animals: prevention by FGL, a peptide mimetic of neural cell adhesion molecule. Behav Pharmacol 19, 4149.Google Scholar
10. Trouche, S, Bontempi, B, Roullet, P, et al. (2009) Recruitment of adult-generated neurons into functional hippocampal networks contributes to updating and strengthening of spatial memory. Proc Natl Acad Sci U S A 106, 59195924.Google Scholar
11. Encinas, JM, Vaahtokari, A & Enikolopov, G (2006) Fluoxetine targets early progenitor cells in the adult brain. Proc Natl Acad Sci U S A 103, 82338238.Google Scholar
12. Lee, HJ, Kim, JW, Yim, SV, et al. (2001) Fluoxetine enhances cell proliferation and prevents apoptosis in dentate gyrus of maternally separated rats. Mol Psychiatry 6, 725728.Google Scholar
13. Sanai, N, Nguyen, T, Ihrie, RA, et al. (2011) Corridors of migrating neurons in the human brain and their decline during infancy. Nature 478, 382386.Google Scholar
14. Wang, C, Liu, F, Liu, YY, et al. (2011) Identification and characterization of neuroblasts in the subventricular zone and rostral migratory stream of the adult human brain. Cell Res 21, 15341550.Google Scholar
15. Spalding, KL, Bergmann, O, Alkass, K, et al. (2013) Dynamics of hippocampal neurogenesis in adult humans. Cell 153, 12191227.Google Scholar
16. Jin, K, Sun, Y, Xie, L, et al. (2003) Directed migration of neuronal precursors into the ischemic cerebral cortex and striatum. Mol Cell Neurosci 24, 171189.Google Scholar
17. Liu, J, Solway, K, Messing, RO, et al. (1998) Increased neurogenesis in the dentate gyrus after transient global ischemia in gerbils. J Neurosci 18, 77687778.Google Scholar
18. Tan, YF, Preston, E & Wojtowicz, JM (2010) Enhanced post-ischemic neurogenesis in aging rats. Front Neurosci 4, 163.Google Scholar
19. Kelly, S, Bliss, TM, Shah, AK, et al. (2004) Transplanted human fetal neural stem cells survive, migrate, and differentiate in ischemic rat cerebral cortex. Proc Natl Acad Sci U S A 101, 1183911844.Google Scholar
20. Chintawar, S, Hourez, R, Ravella, A, et al. (2009) Grafting neural precursor cells promotes functional recovery in an SCA1 mouse model. J Neurosci 29, 1312613135.Google Scholar
21. Cummings, BJ, Uchida, N, Tamaki, SJ, et al. (2005) Human neural stem cells differentiate and promote locomotor recovery in spinal cord-injured mice. Proc Natl Acad Sci U S A 102, 1406914074.Google Scholar
22. Nori, S, Okada, Y, Yasuda, A, et al. (2011) Grafted human-induced pluripotent stem-cell-derived neurospheres promote motor functional recovery after spinal cord injury in mice. Proc Natl Acad Sci U S A 108, 1682516830.CrossRefGoogle ScholarPubMed
23. Xu, Y, Tamamaki, N, Noda, T, et al. (2005) Neurogenesis in the ependymal layer of the adult rat 3rd ventricle. Exp Neurol 192, 251264.Google Scholar
24. Rodriguez, EM, Blazquez, JL, Pastor, FE, et al. (2005) Hypothalamic tanycytes: a key component of brain–endocrine interaction. Int Rev Cytol 247, 89164.Google Scholar
25. Lee, DA & Blackshaw, S (2012) Functional implications of hypothalamic neurogenesis in the adult mammalian brain. Int J Dev Neurosci 30, 615621.Google Scholar
26. McNay, DE, Briancon, N, Kokoeva, MV, et al. (2012) Remodeling of the arcuate nucleus energy-balance circuit is inhibited in obese mice. J Clin Invest 122, 142152.CrossRefGoogle ScholarPubMed
27. Padilla, SL, Carmody, JS & Zeltser, LM (2010) Pomc-expressing progenitors give rise to antagonistic neuronal populations in hypothalamic feeding circuits. Nat Med 16, 403405.CrossRefGoogle ScholarPubMed
28. Li, J, Tang, Y & Cai, D (2012) IKKβ/NF-κB disrupts adult hypothalamic neural stem cells to mediate a neurodegenerative mechanism of dietary obesity and pre-diabetes. Nat Cell Biol 14, 9991012.Google Scholar
29. Pierce, AA & Xu, AW (2010) De novo neurogenesis in adult hypothalamus as a compensatory mechanism to regulate energy balance. J Neurosci 30, 723730.Google Scholar
30. Barry, DS, Pakan, JM & McDermott, KW (2014) Radial glial cells: key organisers in CNS development. Int J Biochem Cell Biol 46, 7679.Google Scholar
31. Ihrie, RA & Alvarez-Buylla, A (2008) Cells in the astroglial lineage are neural stem cells. Cell Tissue Res 331, 179191.Google Scholar
32. Pinto, L & Gotz, M (2007) Radial glial cell heterogeneity – the source of diverse progeny in the CNS. Prog Neurobiol 83, 223.Google Scholar
33. Bergami, M, Rimondini, R, Santi, S, et al. (2008) Deletion of TrkB in adult progenitors alters newborn neuron integration into hippocampal circuits and increases anxiety-like behavior. Proc Natl Acad Sci U S A 105, 1557015575.Google Scholar
34. Sairanen, M, Lucas, G, Ernfors, P, et al. (2005) Brain-derived neurotrophic factor and antidepressant drugs have different but coordinated effects on neuronal turnover, proliferation, and survival in the adult dentate gyrus. J Neurosci 25, 10891094.Google Scholar
35. Li, R & Shen, Y (2005) Estrogen and brain: synthesis, function and diseases. Front Biosci 10, 257267.Google Scholar
36. Sicard, F, Ehrhart-Bornstein, M, Corbeil, D, et al. (2007) Age-dependent regulation of chromaffin cell proliferation by growth factors, dehydroepiandrosterone (DHEA), and DHEA sulfate. Proc Natl Acad Sci U S A 104, 20072012.Google Scholar
37. Denis, I, Potier, B, Vancassel, S, et al. (2013) Omega-3 fatty acids and brain resistance to ageing and stress: body of evidence and possible mechanisms. Ageing Res Rev 12, 579594.Google Scholar
38. Garza, JC, Guo, M, Zhang, W, et al. (2008) Leptin increases adult hippocampal neurogenesis in vivo and in vitro . J Biol Chem 283, 1823818247.CrossRefGoogle ScholarPubMed
39. Han, J, Wang, B, Xiao, Z, et al. (2008) Mammalian target of rapamycin (mTOR) is involved in the neuronal differentiation of neural progenitors induced by insulin. Mol Cell Neurosci 39, 118124.Google Scholar
40. Li, E, Kim, Y, Kim, S, et al. (2014) Ghrelin stimulates proliferation, migration and differentiation of neural progenitors from the subventricular zone in the adult mice. Exp Neurol 252, 7584.Google Scholar
41. Stanic, D, Paratcha, G, Ledda, F, et al. (2008) Peptidergic influences on proliferation, migration, and placement of neural progenitors in the adult mouse forebrain. Proc Natl Acad Sci U S A 105, 36103615.Google Scholar
42. Parthsarathy, V & Holscher, C (2013) Chronic treatment with the GLP1 analogue liraglutide increases cell proliferation and differentiation into neurons in an AD mouse model. PLOS ONE 8, e58784.Google Scholar
43. Larsen, C & Grattan, D (2010) Prolactin-induced mitogenesis in the subventricular zone of the maternal brain during early pregnancy is essential for normal postpartum behavorial responses in the mother. Endocrinology 151, 38053814.CrossRefGoogle Scholar
44. Galvao, RP, Garcia-Verdugo, JM & Alvarez-Buylla, A (2008) Brain-derived neurotrophic factor signaling does not stimulate subventricular zone neurogenesis in adult mice and rats. J Neurosci 28, 1336813383.Google Scholar
45. Bondolfi, L, Ermini, F, Long, JM, et al. (2004) Impact of age and caloric restriction on neurogenesis in the dentate gyrus of C57BL/6 mice. Neurobiol Aging 25, 333340.Google Scholar
46. Kuhn, HG, Dickinson-Anson, H & Gage, FH (1996) Neurogenesis in the dentate gyrus of the adult rat: age-related decrease of neuronal progenitor proliferation. J Neurosci 16, 20272033.Google Scholar
47. Zhang, G, Li, J, Purkayastha, S, et al. (2013) Hypothalamic programming of systemic ageing involving IKK-β, NF-κB and GnRH. Nature 497, 211216.Google Scholar
48. Aoki, H, Kimoto, K, Hori, N, et al. (2005) Cell proliferation in the dentate gyrus of rat hippocampus is inhibited by soft diet feeding. Gerontology 51, 369374.Google Scholar
49. Nose-Ishibashi, K, Watahiki, J, Yamada, K, et al. (2014) Soft-diet feeding after weaning affects behavior in mice: potential increase in vulnerability to mental disorders. Neuroscience 263, 257268.Google Scholar
50. Canto, C & Auwerx, J (2009) Caloric restriction, SIRT1 and longevity. Trends Endocrinol Metab 20, 325331.Google Scholar
51. Pijl, H (2012) Longevity. The allostatic load of dietary restriction. Physiol Behav 106, 5157.Google Scholar
52. Ruetenik, A & Barrientos, A (2015) Dietary restriction, mitochondrial function and aging: from yeast to humans. Biochim Biophys Acta 1847, 14341447.Google Scholar
53. Lee, J, Duan, W, Long, JM, et al. (2000) Dietary restriction increases the number of newly generated neural cells, and induces BDNF expression, in the dentate gyrus of rats. J Mol Neurosci 15, 99108.Google Scholar
54. Lee, J, Duan, W & Mattson, MP (2002) Evidence that brain-derived neurotrophic factor is required for basal neurogenesis and mediates, in part, the enhancement of neurogenesis by dietary restriction in the hippocampus of adult mice. J Neurochem 82, 13671375.Google Scholar
55. Van der Borght, K, Köhnke, R, Göransson, N, et al. (2011) Reduced neurogenesis in the rat hippocampus following high fructose consumption. Regul Pept 167, 2630.Google Scholar
56. Lindqvist, A, Mohapel, P, Bouter, B, et al. (2006) High-fat diet impairs hippocampal neurogenesis in male rats. Eur J Neurol 13, 13851388.Google Scholar
57. Rivera, P, Romero-Zerbo, Y, Pavon, FJ, et al. (2011) Obesity-dependent cannabinoid modulation of proliferation in adult neurogenic regions. Eur J Neurosci 33, 15771586.Google Scholar
58. Yon, MA, Mauger, SL & Pickavance, LC (2013) Relationships between dietary macronutrients and adult neurogenesis in the regulation of energy metabolism. Br J Nutr 109, 15731589.Google Scholar
59. Lee, D, Bedont, J, Pak, T, et al. (2012) Tanycytes of the hypothalamic median eminence form a diet-responsive neurogenic niche. in Nat Neurosci 15, 700702.Google Scholar
60. Lee, DA, Yoo, S, Pak, T, et al. (2014) Dietary and sex-specific factors regulate hypothalamic neurogenesis in young adult mice. Front Neurosci 8, 157.Google Scholar
61. Tozuka, Y, Kumon, M, Wada, E, et al. (2010) Maternal obesity impairs hippocampal BDNF production and spatial learning performance in young mouse offspring. Neurochem Int 57, 235247.Google Scholar
62. Jacka, FN, Cherbuin, N, Anstey, KJ, et al. (2015) Western diet is associated with a smaller hippocampus: a longitudinal investigation. BMC Med 13, 215.Google Scholar
63. Rendeiro, C, Guerreiro, JD, Williams, CM, et al. (2012) Flavonoids as modulators of memory and learning: molecular interactions resulting in behavioural effects. Proc Nutr Soc 71, 246262.Google Scholar
64. Lee, S, Kim, DH, Lee, DH, et al. (2010) Oroxylin A, a flavonoid, stimulates adult neurogenesis in the hippocampal dentate gyrus region of mice. Neurochem Res 35, 17251732.CrossRefGoogle ScholarPubMed
65. Okuyama, S, Minami, S, Shimada, N, et al. (2013) Anti-inflammatory and neuroprotective effects of auraptene, a citrus coumarin, following cerebral global ischemia in mice. Eur J Pharmacol 699, 118123.CrossRefGoogle ScholarPubMed
66. Okuyama, S, Shimada, N, Kaji, M, et al. (2012) Heptamethoxyflavone, a citrus flavonoid, enhances brain-derived neurotrophic factor production and neurogenesis in the hippocampus following cerebral global ischemia in mice. Neurosci Lett 528, 190195.Google Scholar
67. Oh, SB, Park, HR, Jang, YJ, et al. (2013) Baicalein attenuates impaired hippocampal neurogenesis and the neurocognitive deficits induced by γ-ray radiation. Br J Pharmacol 168, 421431.Google Scholar
68. Nakajima, K, Niisato, N & Marunaka, Y (2011) Quercetin stimulates NGF-induced neurite outgrowth in PC12 cells via activation of Na+/K+/2Cl cotransporter. Cell Physiol Biochem 28, 147156.Google Scholar
69. Oberbauer, E, Urmann, C, Steffenhagen, C, et al. (2013) Chroman-like cyclic prenylflavonoids promote neuronal differentiation and neurite outgrowth and are neuroprotective. J Nutr Biochem 24, 19531962.CrossRefGoogle ScholarPubMed
70. Madhyastha, S, Sekhar, S & Rao, G (2013) Resveratrol improves postnatal hippocampal neurogenesis and brain derived neurotrophic factor in prenatally stressed rats. Int J Dev Neurosci 31, 580585.Google Scholar
71. Moriya, J, Chen, R, Yamakawa, J, et al. (2011) Resveratrol improves hippocampal atrophy in chronic fatigue mice by enhancing neurogenesis and inhibiting apoptosis of granular cells. Biol Pharm Bull 34, 354359.Google Scholar
72. Kim, SJ, Son, TG, Park, HR, et al. (2008) Curcumin stimulates proliferation of embryonic neural progenitor cells and neurogenesis in the adult hippocampus. J Biol Chem 283, 1449714505.Google Scholar
73. Dong, S, Zeng, Q, Mitchell, ES, et al. (2012) Curcumin enhances neurogenesis and cognition in aged rats: implications for transcriptional interactions related to growth and synaptic plasticity. PLOS ONE 7, e31211.Google Scholar
74. Wang, YF, Zu, JN, Li, J, et al. (2014) Curcumin promotes the spinal cord repair via inhibition of glial scar formation and inflammation. Neurosci Lett 560, 5156.CrossRefGoogle ScholarPubMed
75. Coti Bertrand, P, O’Kusky, JR & Innis, SM (2006) Maternal dietary (n-3) fatty acid deficiency alters neurogenesis in the embryonic rat brain. J Nutr 136, 15701575.Google Scholar
76. Latour, A, Grintal, B, Champeil-Potokar, G, et al. (2013) Omega-3 fatty acids deficiency aggravates glutamatergic synapse and astroglial aging in the rat hippocampal CA1. Aging Cell 12, 7684.Google Scholar
77. Calderon, F & Kim, HY (2004) Docosahexaenoic acid promotes neurite growth in hippocampal neurons. J Neurochem 90, 979988.Google Scholar
78. Cao, D, Kevala, K, Kim, J, et al. (2009) Docosahexaenoic acid promotes hippocampal neuronal development and synaptic function. J Neurochem 111, 510521.CrossRefGoogle ScholarPubMed
79. He, C, Qu, X, Cui, L, et al. (2009) Improved spatial learning performance of fat-1 mice is associated with enhanced neurogenesis and neuritogenesis by docosahexaenoic acid. Proc Natl Acad Sci U S A 106, 1137011375.Google Scholar
80. Dyall, SC, Michael, GJ & Michael-Titus, AT (2010) Omega-3 fatty acids reverse age-related decreases in nuclear receptors and increase neurogenesis in old rats. J Neurosci Res 88, 20912102.Google Scholar
81. Cutuli, D, De Bartolo, P, Caporali, P, et al. (2014) n-3 Polyunsaturated fatty acids supplementation enhances hippocampal functionality in aged mice. Front Aging Neurosci 6, 220.Google Scholar
82. Goustard-Langelier, B, Koch, M, Lavialle, M, et al. (2013) Rat neural stem cell proliferation and differentiation are durably altered by the in utero polyunsaturated fatty acid supply. J Nutr Biochem 24, 380387.Google Scholar
83. Andrade, JP, Mesquita, R, Assuncao, M, et al. (2006) Effects of food restriction on synthesis and expression of brain-derived neurotrophic factor and tyrosine kinase B in dentate gyrus granule cells of adult rats. Neurosci Lett 399, 135140.Google Scholar
84. Gray, J, Yeo, GS, Cox, JJ, et al. (2006) Hyperphagia, severe obesity, impaired cognitive function, and hyperactivity associated with functional loss of one copy of the brain-derived neurotrophic factor (BDNF) gene. Diabetes 55, 33663371.Google Scholar
85. Jeon, SJ, Rhee, SY, Seo, JE, et al. (2011) Oroxylin A increases BDNF production by activation of MAPK-CREB pathway in rat primary cortical neuronal culture. Neurosci Res 69, 214222.CrossRefGoogle ScholarPubMed
86. Rong, W, Wang, J, Liu, X, et al. (2012) Naringin treatment improves functional recovery by increasing BDNF and VEGF expression, inhibiting neuronal apoptosis after spinal cord injury. Neurochem Res 37, 16151623.Google Scholar
87. Williams, CM, El Mohsen, MA, Vauzour, D, et al. (2008) Blueberry-induced changes in spatial working memory correlate with changes in hippocampal CREB phosphorylation and brain-derived neurotrophic factor (BDNF) levels. Free Radic Biol Med 45, 295305.Google Scholar
88. Zhang, F, Lu, YF, Wu, Q, et al. (2012) Resveratrol promotes neurotrophic factor release from astroglia. Exp Biol Med 237, 943948.CrossRefGoogle ScholarPubMed
89. Rahvar, M, Nikseresht, M, Shafiee, SM, et al. (2011) Effect of oral resveratrol on the BDNF gene expression in the hippocampus of the rat brain. Neurochem Res 36, 761765.Google Scholar
90. Liu, D, Xie, K, Yang, X, et al. (2014) Resveratrol reverses the effects of chronic unpredictable mild stress on behavior, serum corticosterone levels and BDNF expression in rats. Behavl Brain Res 264, 916.Google Scholar
91. Wang, Z, Gu, J, Wang, X, et al. (2013) Antidepressant-like activity of resveratrol treatment in the forced swim test and tail suspension test in mice: the HPA axis, BDNF expression and phosphorylation of ERK. Pharmacol Biochem Behav 112, 104110.Google Scholar
92. Wu, A, Ying, Z & Gomez-Pinilla, F (2004) Dietary omega-3 fatty acids normalize BDNF levels, reduce oxidative damage, and counteract learning disability after traumatic brain injury in rats. J Neurotrauma 21, 14571467.Google Scholar
93. Sable, PS, Kale, AA & Joshi, SR (2013) Prenatal omega 3 fatty acid supplementation to a micronutrient imbalanced diet protects brain neurotrophins in both the cortex and hippocampus in the adult rat offspring. Metabolism 62, 16071622.Google Scholar
94. Finkbeiner, S (2000) CREB couples neurotrophin signals to survival messages. Neuron 25, 1114.Google Scholar
95. Wu, A, Ying, Z & Gomez-Pinilla, F (2007) Omega-3 fatty acids supplementation restores mechanisms that maintain brain homeostasis in traumatic brain injury. J Neurotrauma 24, 15871595.Google Scholar
96. Kumar, TP, Antony, S, Gireesh, G, et al. (2010) Curcumin modulates dopaminergic receptor, CREB and phospholipase C gene expression in the cerebral cortex and cerebellum of streptozotocin induced diabetic rats. J Biomed Sci 17, 43.Google Scholar
97. Xu, Y, Ku, B, Tie, L, et al. (2006) Curcumin reverses the effects of chronic stress on behavior, the HPA axis, BDNF expression and phosphorylation of CREB. Brain Res 1122, 5664.Google Scholar
98. Lin, Y, Chen, F, Zhang, J, et al. (2013) Neuroprotective effect of resveratrol on ischemia/reperfusion injury in rats through TRPC6/CREB pathways. J Mol Neurosci 50, 504513.Google Scholar
99. Tang, Y, Purkayastha, S & Cai, D (2015) Hypothalamic microinflammation: a common basis of metabolic syndrome and aging. Trends Neurosci 38, 3644.Google Scholar
100. Chaudhari, N, Talwar, P, Parimisetty, A, et al. (2014) A molecular web: endoplasmic reticulum stress, inflammation, and oxidative stress. Front Cell Neurosci 8, 213.Google Scholar
101. Baker, RG, Hayden, MS & Ghosh, S (2011) NF-κB, inflammation, and metabolic disease. Cell Metab 13, 1122.Google Scholar
102. Yeung, F, Hoberg, JE, Ramsey, CS, et al. (2004) Modulation of NF-κB-dependent transcription and cell survival by the SIRT1 deacetylase. EMBO J 23, 23692380.Google Scholar
103. Pfluger, PT, Herranz, D, Velasco-Miguel, S, et al. (2008) Sirt1 protects against high-fat diet-induced metabolic damage. Proc Natl Acad Sci U S A 105, 97939798.Google Scholar
104. Howitz, KT, Bitterman, KJ, Cohen, HY, et al. (2003) Small molecule activators of sirtuins extend Saccharomyces cerevisiae lifespan. Nature 425, 191196.Google Scholar
105. Wood, JG, Rogina, B, Lavu, S, et al. (2004) Sirtuin activators mimic caloric restriction and delay ageing in metazoans. Nature 430, 686689.Google Scholar
106. Davis, JM, Murphy, EA, Carmichael, MD, et al. (2009) Quercetin increases brain and muscle mitochondrial biogenesis and exercise tolerance. Am J Physiol Regul Integr Comp Physiol 296, R1071R1077.Google Scholar
107. Su, KY, Yu, CY, Chen, YW, et al. (2014) Rutin, a flavonoid and principal component of saussurea involucrata, attenuates physical fatigue in a forced swimming mouse model. Int J Med Sci 11, 528537.Google Scholar
108. Prozorovski, T, Schulze-Topphoff, U, Glumm, R, et al. (2008) Sirt1 contributes critically to the redox-dependent fate of neural progenitors. Nat Cell Biol 10, 385394.CrossRefGoogle Scholar
109. Michán, S, Li, Y, Chou, MM, et al. (2010) SIRT1 is essential for normal cognitive function and synaptic plasticity. J Neurosci 30, 96959707.Google Scholar
110. Zocchi, L & Sassone-Corsi, P (2012) SIRT1-mediated deacetylation of MeCP2 contributes to BDNF expression. Epigenetics 7, 695700.Google Scholar
111. Jiang, M, Wang, J, Fu, J, et al. (2012) Neuroprotective role of Sirt1 in mammalian models of Huntington’s disease through activation of multiple Sirt1 targets. Nat Med 18, 153158.Google Scholar
112. Ruderman, NB, Xu, XJ, Nelson, L, et al. (2010) AMPK and SIRT1: a long-standing partnership? Am J Physiol Endocrinol Metab 298, E751E760.Google Scholar
113. Dasgupta, B & Milbrandt, J (2009) AMP-activated protein kinase phosphorylates retinoblastoma protein to control mammalian brain development. Dev Cell 16, 256270.Google Scholar
114. Eid, HM, Martineau, LC, Saleem, A, et al. (2010) Stimulation of AMP-activated protein kinase and enhancement of basal glucose uptake in muscle cells by quercetin and quercetin glycosides, active principles of the antidiabetic medicinal plant Vaccinium vitis-idaea . Mol Nutr Food Res 54, 9911003.Google Scholar
115. Lu, J, Wu, DM, Zheng, YL, et al. (2010) Quercetin activates AMP-activated protein kinase by reducing PP2C expression protecting old mouse brain against high cholesterol-induced neurotoxicity. J Pathol 222, 199212.Google Scholar
116. Tong, X, Smith, KA & Pelling, JC (2012) Apigenin, a chemopreventive bioflavonoid, induces AMP-activated protein kinase activation in human keratinocytes. Mol Carcinog 51, 268279.Google Scholar
117. Zygmunt, K, Faubert, B, MacNeil, J, et al. (2010) Naringenin, a citrus flavonoid, increases muscle cell glucose uptake via AMPK. Biochem Biophys Res Commun 398, 178183.Google Scholar
118. Vingtdeux, V, Giliberto, L, Zhao, H, et al. (2010) AMP-activated protein kinase signaling activation by resveratrol modulates amyloid-β peptide metabolism. J Biol Chem 285, 91009113.Google Scholar
119. Jing, H, Yao, J, Liu, X, et al. (2014) Fish-oil emulsion (omega-3 polyunsaturated fatty acids) attenuates acute lung injury induced by intestinal ischemia–reperfusion through adenosine 5’-monophosphate-activated protein kinase–sirtuin1 pathway. J Surg Res 187, 252261.Google Scholar
120. Xue, B, Yang, Z, Wang, X, et al. (2012) Omega-3 polyunsaturated fatty acids antagonize macrophage inflammation via activation of AMPK/SIRT1 pathway. PLOS ONE 7, e45990.Google Scholar
121. Witte, AV, Kerti, L, Margulies, DS, et al. (2014) Effects of resveratrol on memory performance, hippocampal functional connectivity, and glucose metabolism in healthy older adults. J Neurosci 34, 78627870.Google Scholar
122. Huhn, S, Kharabian Masouleh, S, Stumvoll, M, et al. (2015) Components of a Mediterranean diet and their impact on cognitive functions in aging. Front Aging Neurosci 7, 132.Google Scholar
123. Diaz Heijtz, R, Wang, S, Anuar, F, et al. (2011) Normal gut microbiota modulates brain development and behavior. Proc Natl Acad Sci U S A 108, 30473052.Google Scholar
124. Crumeyrolle-Arias, M, Jaglin, M, Bruneau, A, et al. (2014) Absence of the gut microbiota enhances anxiety-like behavior and neuroendocrine response to acute stress in rats. Psychoneuroendocrinology 42, 207217.Google Scholar
125. Clarke, G, Grenham, S, Scully, P, et al. (2013) The microbiome–gut–brain axis during early life regulates the hippocampal serotonergic system in a sex-dependent manner. Mol Psychiatry 18, 666673.Google Scholar
126. Tremaroli, V & Backhed, F (2012) Functional interactions between the gut microbiota and host metabolism. Nature 489, 242249.Google Scholar
127. Turnbaugh, PJ, Hamady, M, Yatsunenko, T, et al. (2009) A core gut microbiome in obese and lean twins. Nature 457, 480484.Google Scholar
128. Alang, N & Kelly, CR (2015) Weight gain after fecal microbiota transplantation. Open Forum Infect Dis 2, ofv004.Google Scholar
129. Song, J, Kang, SM, Kim, E, et al. (2015) Adiponectin receptor-mediated signaling ameliorates cerebral cell damage and regulates the neurogenesis of neural stem cells at high glucose concentrations: an in vivo and in vitro study. Cell Death Dis 6, e1844.Google Scholar
130. Fernandez-Real, JM, Serino, M, Blasco, G, et al. (2015) Gut microbiota interacts with brain microstructure and function. J Clin Endocrinol Metab 100, 45054513.Google Scholar
131. Zhu, Y, Lin, X, Zhao, F, et al. (2015) Meat, dairy and plant proteins alter bacterial composition of rat gut bacteria. Sci Rep 5, 15220.Google Scholar
Figure 0

Fig. 1 Adult neurogenesis is a flexible process, which can be modified by positive or negative influences. Pro-inflammatory molecules or mechanisms activate NF-κB, which induces the transcription of pro-inflammatory cytokines, while anti-inflammatory molecules by their action on sirtuin 1 (Sirt1) activate cAMP response element binding protein (CREB), brain-derived neurotrophic factor (BDNF) and AMP-activated protein kinase (AMPK) and restore cell renewal in the brain. The influence of gut microbiota is here hypothetical. For a colour figure, see the online version.