Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Brain-derived neurotrophic factor (BDNF) mediates energy metabolism and feeding behavior. As a neurotrophin, BDNF promotes neuronal differentiation, survival during early development, adult neurogenesis, and neural plasticity; thus, there is the potential that BDNF could modify circuits important to eating behavior and energy expenditure. The possibility that "faulty" circuits could be remodeled by BDNF is an exciting concept for new therapies for obesity and eating disorders. In the hypothalamus, BDNF and its receptor, tropomyosin-related kinase B (TrkB), are extensively expressed in areas associated with feeding and metabolism. Hypothalamic BDNF and TrkB appear to inhibit food intake and increase energy expenditure, leading to negative energy balance. In the hippocampus, the involvement of BDNF in neural plasticity and neurogenesis is important to learning and memory, but less is known about how BDNF participates in energy homeostasis. We review current research about BDNF in specific brain locations related to energy balance, environmental, and behavioral influences on BDNF expression and the possibility that BDNF may influence energy homeostasis via its role in neurogenesis and neural plasticity.

Free full text 


Logo of ajpreguLink to Publisher's site
Am J Physiol Regul Integr Comp Physiol. 2011 May; 300(5): R1053–R1069.
Published online 2011 Feb 23. https://doi.org/10.1152/ajpregu.00776.2010
PMCID: PMC3293512
PMID: 21346243

The lighter side of BDNF

Abstract

Brain-derived neurotrophic factor (BDNF) mediates energy metabolism and feeding behavior. As a neurotrophin, BDNF promotes neuronal differentiation, survival during early development, adult neurogenesis, and neural plasticity; thus, there is the potential that BDNF could modify circuits important to eating behavior and energy expenditure. The possibility that “faulty” circuits could be remodeled by BDNF is an exciting concept for new therapies for obesity and eating disorders. In the hypothalamus, BDNF and its receptor, tropomyosin-related kinase B (TrkB), are extensively expressed in areas associated with feeding and metabolism. Hypothalamic BDNF and TrkB appear to inhibit food intake and increase energy expenditure, leading to negative energy balance. In the hippocampus, the involvement of BDNF in neural plasticity and neurogenesis is important to learning and memory, but less is known about how BDNF participates in energy homeostasis. We review current research about BDNF in specific brain locations related to energy balance, environmental, and behavioral influences on BDNF expression and the possibility that BDNF may influence energy homeostasis via its role in neurogenesis and neural plasticity.

Keywords: food intake, body weight, ventromedial hypothalamus, paraventricular nucleus, brain-derived neurotrophic factor

brain-derived neurotrophic factor (BDNF) is a member of the neurotrophin family of growth factors (151), along with nerve growth factor (152), and neurotrophin (NT) 3 (67, 163), NT 4/5 (28), and NT 6 (88). Neurotrophins are synthesized as 32–35-kDa pro-isoforms, which are later cleaved to mature forms that dimerize after translation and then act as receptor ligands (136). Whereas the precursor forms of other neurotrophins are constitutively secreted, the 32-kDa pro-BDNF is packaged into vesicles of a regulated pathway and is secreted in an activity-dependent manner (87). Pro-BDNF may be secreted as is (48), cleaved by the extracellular protease plasmin (202), or interact with the pan-neurotrophin receptor p75NTR and other receptors that cause an independent biological effect (244). Alternatively, pro-BDNF is processed to the mature form intracellularly by furin or proconvertases, where it forms C-terminal dimers (212, 226).

Mature BDNF is considered the biologically active form, which has a high affinity for the tropomyosin-related kinase B (TrkB) receptor (130). Both BDNF and TrkB are present in presynaptic axon terminals and postsynaptic dendritic compartments of neurons, and they are capable of bidirectional release and activity [for review, see Tyler et al. (259)]. Typical of the neurotrophic factors, BDNF stimulates the development and differentiation of new neurons (3, 131) and promotes long-term potentiation (LTP) (139, 140, 205), and neuron survival (97, 105, 116). BDNF is abundantly expressed throughout the developing and mature CNS and in many peripheral tissues, including muscle, liver, and adipose (42, 159, 182, 261). Regional differences between BDNF mRNA levels and protein concentrations in the CNS are often reported (7, 51, 192, 193), which may be related to regulatory mechanisms, mRNA decay (164), or BDNF anterograde transport (7).

BDNF is synthesized in several areas of the hypothalamus, including the paraventricular nucleus (PVN), the ventromedial hypothalamic nucleus (VMN), the dorsomedial hypothalamic nucleus (DMN), and the lateral hypothalamic area (LH) (51). Additionally, BDNF-immunoreactive fibers have been identified in the arcuate nucleus (Arc); however, the Arc does not appear to be a site of BDNF synthesis (51). BDNF is also widely expressed throughout the hippocampus; amygdala; select areas of the thalamus, including the ventral tegmental area (VTA); and in areas of the hindbrain, including the dorsal vagal complex (DVC) (17, 51). In recent years, much attention has been given to BDNF for its role in energy homeostasis. While examining BDNF and neuronal plasticity in vivo, Lapchak et al. (145) noted that chronic intraventricular administration of BDNF prevented weight gain. It has since been observed in many studies that central administration of BDNF induces appetite suppression and weight loss (207, 267, 269, 270), increases locomotor activity (186), and resting metabolic rate (266, 268). An obese phenotype is also observed in BDNF-conditional knockout mice, where BDNF is deleted after birth and the knockout is restricted to the brain (216). In addition to central effects, BDNF exerts peripheral actions that affect glucose metabolism (287, 289, 290), energy expenditure (288, 290), and food intake (287). Both central and peripherally administered BDNF lowers blood glucose and increases energy expenditure in animal models of type 2 diabetes (188). The combined effects of central and peripheral BDNF are apparent from instances where BDNF is globally reduced, as is the case in rodents and human subjects with haploinsufficiency for the gene, which results in obesity and hyperphagia (91, 99, 123). The neurotrophin receptor TrkB is a receptor tyrosine kinase, which upon activation, results in receptor dimerization, followed by receptor transphosphorylation and the initiation of intracellular signaling cascades. A human mutation affecting the ability of the TrkB receptor to autophosphorylate is associated with obesity and hyperphagia (291). The TrkB receptor exists in full length and two truncated forms, but only the full-length receptor contains intracellular tyrosine kinase activity (4). The two truncated forms are generated by alternative splicing of the full-length receptor, and they are capable of inhibiting activity of BDNF by forming heterodimers with full-length TrkB (65), or by binding and internalizing BDNF (98). The role of BDNF in obesity appears to, at least in part, involve signaling through the full-length TrkB receptor, as TrkB hypomorphs, expressing ¼ of the full-length TrkB receptor of wild-type mice, are obese (283, 284). In the absence of BDNF the antiobesity effect of TrkB is still possible. This is evidenced by the use of two different BDNF agonists, the TrkB ligand NT4 and a TrkB-specific antibody that acts as a receptor agonist, which when administered to the hypothalamus of mice, caused reductions in food intake and resistance to diet-induced obesity (DIO), as well as polygenic and leptin receptor deficiency-associated obesity (255). It is important to note that BDNF appears to be the main natural ligand for TrkB, as rodents heterozygous for BDNF, but not NT4, display the obese phenotype (123). Thus, both BDNF and TrkB are necessary for BDNF-mediated effects on energy balance (255).

The human BDNF gene contains a total of 10 exons coding for the 5′ untranslated region and are alternatively spliced to a common 3′ coding exon, resulting in 34 possible mRNA transcripts (211). In rodents BDNF contains 9 exons, encoding for 24 different mRNA transcripts, each of them ultimately translating into an identical mature BDNF [reviewed by Cunha (54)]. The expression of BDNF transcripts is tissue-specific, that is, it is differentially expressed throughout different brain sites and peripheral tissues (29, 247). Moreover, environmental cues, such as stress, can alter transcript expression (80, 167, 248), which is associated with alterations in pro-BDNF/total BDNF ratio (248). The exact role of each of these individual transcripts is still largely unknown. In the VMN of rodents, transcripts encoding for BDNF exons I, II, and IV are expressed, whereas exon III is not (254). Transcription of exons I and IV in this region is regulated by steroidogenic factor 1 (SF-1), which, when reduced (as in the case of SF-1 heterozygotes), impairs hypothalamic function and results in hyperphagia and weight gain (254). Different genetic variants impact the activity of BDNF by affecting biosynthesis and/or post-translational processing of the pro-BDNF precursor (183). One variant, in particular, the single nucleotide polymorphism (SNP), which results in a substitution of a valine for methionine residue at position 66 (Val66Met), has been identified as having a strong correlation with eating disorders (ED), specifically anorexia nervosa (AN) of the restricting type, and is associated with low body mass index (BMI) (214). This polymorphism alters the intracellular packaging of pro-BDNF and may affect activity-dependent secretion of the mature BDNF peptide (48, 64). According to a recent meta-analysis, the Val66Met polymorphism is associated with a 33% increased risk for ED (90); however, this analysis did not subdivide ED by category. Recent evidence is conflicting regarding this gene variant, as some have found no association between ED and Val66Met (11), or AN (56), and one study reported a link between Val66Met and obesity in females (24). The conflicting evidence regarding ED and the BDNF SNP Val66Met reflects both the complexity of eating disorders and the range of factors affecting feeding behavior.

In a recent genome-wide association study, BDNF was one of 18 gene loci, where having a certain SNP variant was associated with higher BMI (235). Another study examined 41 different SNPs near the BDNF locus in 87 adults with sudden death and made comparisons between BMI and BDNF mRNA in the VMN of the cadavers. In the subjects with extreme obesity (BMI ≥ 40 kg/m2), BDNF expression was reduced twofold compared with overweight and obese individuals. There was an association between homozygosity for the minor C allele at rs12291063, reduced VMN BDNF expression, and high BMI, which suggests that having the SNP at rs12291063 may be a risk factor for obesity (201). In mice, heterozygosity for Bdnf decreases hypothalamic expression and results in hyperphagia and obesity (123). WAGR (Wilms' tumor, aniridia, genitourinary anomalies, and mental retardation) syndrome is a rare disorder characterized by heterozygous gene deletions in at least two genes located near BDNF in the 11p13 region, and sometimes accompanies the heterozygous deletion of BDNF. In a study of individuals with WAGR syndrome, 100% of those heterozygous for BDNF deletion were obese by the age of 10, in contrast with 20% of those without BDNF deletion (99).

BDNF and the Central Regulation of Energy Metabolism

BDNF was first observed to affect energy metabolism with intracerebroventricular administration (207). In recent years, studies have identified additional sites of BDNF action regulating energy balance, including the DVC, hypothalamic PVN and VMN, the VTA, amygdala, and possibly the hippocampus (17, 18, 32, 52, 57, 58, 266270, 272). In regulating energy balance, BDNF interacts with several other neuropeptides, including melanocortin (18, 39, 196, 255, 284), leptin (17, 18, 40, 137, 268, 269, 272), corticotrophin-releasing hormone (CRH) (35, 39, 40, 251, 271), and thyrotropin releasing hormone (TRH) (35, 233, 260).

In 1995, Pelleymounter et al. (207) discovered that intracerebroventricular administration of BDNF decreased energy intake and body weight of rats, which was associated with a dose-dependent increase in serotonin turnover. A pair-fed group had comparable weight loss, but the recovery weight gain in these rats was much faster than the BDNF-infused group (207), suggesting BDNF promotes lasting metabolic changes. Additional evidence for a central role of BDNF in the regulation of energy balance is apparent, as animals with reduced Bdnf expression, either due to a conditional homozygous knockout in the brain or due to heterozygous gene expression (Bdnf +/), develop hyperphagia, obesity, and resistance to insulin and leptin (123, 216). Administration of BDNF intracerebroventricularly reverses the hyperphagic and obese phenotype of Bdnf +/mutant mice (123). A single intracerebroventricular injection of BDNF is sufficient to improve insulin receptor signaling in the liver of streptozotocin-induced diabetic mice, whereas no direct effect of BDNF on cultured hepatocytes has been observed (256). This indicates that independently of anorectic effects, centrally administered BDNF may affect glucose metabolism. Nonomura et al. (197) observed that intracerebroventricular BDNF dose-dependently lowers blood glucose and increases pancreatic insulin content in leptin receptor-deficient db/db mice and does so independently of food intake. These improvements were also associated with increased norepinephrine turnover and uncoupling protein-1 (UCP-1) expression in brown adipose tissue (BAT) (197). Taken together, it appears the central BDNF enhances energy expenditure via activation of the sympathetic nervous system and improves blood glucose in obese, diabetic rodents. It should be noted that BDNF does not affect blood glucose in normoglycemic rats (207).

The adult rat hypothalamus contains high levels of BDNF (119, 193, 241), and most hypothalamic neurons express the TrkB receptor (43, 168, 176). Overexpression of the Bdnf gene in the hypothalamus is associated with increased heat production, respiratory exchange ratio, and resting metabolism, and increased hypothalamic expression of TrkB, insulin receptor, CRH, and TRH (39, 40). Increased Bdnf expression is also associated with sharp decreases in leptin and insulin concentrations, and increases in the adipose tissue-secreted hormone adiponectin (39), which is associated with increased fatty acid oxidation, enhanced glucose metabolism, and weight loss (79). We have found that BDNF reduces food intake and influences energy expenditure when injected into certain specific hypothalamic sites, but not others. For example, although the LH expresses BDNF and its receptor, specific site injection of BDNF did not significantly reduce feeding and body weight (267). The hypothalamic PVN is responsive to physiological stimuli, is involved in stress responses (21, 86), and contains high levels of BDNF and TrkB mRNA (241). We found that injections of BDNF in the PVN increases energy expenditure, mainly by increasing resting metabolic rate, and increasing thermogenic capacity, as indicated by elevation of UCP-1 expression in BAT (266). We have also found that a single injection of BDNF reduces food intake and body weight (267) for up to 48 h after injection, suggesting a prolonged and potent effect. Animals who were made obese with a high-fat diet (HFD) and subsequently given PVN injections of BDNF on alternate days over an extended period, had significant reductions in energy intake, body weight, and body fat (including visceral fat), compared with artificial cerebrospinal fluid-injected controls. BDNF also normalized HFD-induced hyperglycemia, hyperlipidemia, hyperinsulinemia, and hyperleptinemia, suggesting chronic BDNF in the PVN improves metabolic syndrome and associated resistance to insulin and leptin (270). Furthermore, the animals more susceptible to DIO were also more sensitive to PVN injections of BDNF (270).

Stress paradigms increase expression of BDNF mRNA in the PVN (213), which is associated with decreased inhibitory synaptic input leading to activation of PVN neurons (264, 265). The mechanism for BDNF-associated removal of inhibition was elucidated by Hewitt and Bains (103), who observed that through TrkB receptor activation on postsynaptic neurons, BDNF reduces the surface expression of inhibitory GABAA receptor clusters in the PVN. Thus, it is likely that BDNF increases the firing rate of PVN neurons involved in the stress response. CRH neurons in the PVN regulate locomotor activity and body temperature (158, 217), and, thus, the removal of inhibitory GABAA receptors on these neurons would likely account for the physiological and behavioral effects observed when BDNF is injected directly in this area. Naert et al. (186) observed that BDNF infused continuously into the lateral ventricle of rats causes a significant increase in paraventricular CRH and AVP mRNA, which is associated with increased locomotor activity, body temperature, and reduced body weights (186). BDNF overexpression in the hypothalamus also resulted in increased CRH expression (39), and hypothalamic increases in BDNF and CRH were associated with β-adrenergic receptor activation (40). Again, these data support the notion that the sympathetic nervous system is important in BDNF-associated metabolic changes. Toriya et al. (251) report that PVN-injected BDNF-induced reduction in feeding and body weight is mediated via CRH-R2, and accordingly, the effects of BDNF were blocked using a CRH antagonist. Similarly, we have observed that the effect of BDNF on feeding and body weight gain due to BDNF injections in the VMN or PVN was attenuated by pretreatment with a CRH antagonist (271). In the PVN, mRNA for BDNF and CRH are colocalized (186), and there is also colocalization for TrkB receptor and CRH (251), suggesting potential signaling between the two neuropeptides in the regulation of energy metabolism. (Fig. 1)

An external file that holds a picture, illustration, etc.
Object name is zh60051175620001.jpg

Brain-derived neurotrophic factor (BDNF) and the central regulation of energy metabolism. Leptin is secreted from adipose tissue and activates receptors of the anorexigenic proopiomelanocortin (POMC) neurons of the arcuate nucleus (Arc) and neurons of the ventromedial nucleus of the hypothalamus (VMN). POMC neurons project to the hypothalamic paraventricular nucleus (PVN) and VMN, where they secrete alpha-melanocyte-stimulating hormone (α-MSH), which binds to the melanocortin receptor (MC4R). MC4R activation in the VMN controls BDNF expression; however, arcuate POMC neurons project sparsely to the VMN, and it is currently unknown whether α-MSH coming from these projections affects BDNF expression (284). Both leptin and α-MSH induce the expression of BDNF in the VMN. BDNF activates the tropomyosin-related kinase B (TrkB) receptor in neurons of the VMN and PVN, with consequences for energy metabolism (boxes on the left). In PVN neurons, BDNF removes inhibitory GABAA receptor clusters, allowing for greater neuronal excitability. Neurons of the PVN express corticotropin-releasing hormone (CRH) and in the PVN, effects of BDNF are attenuated when the CRH receptor is blocked by a receptor antagonist, indicating that BDNF activates the CRH-urocortin-CRH-R2 pathway. BDNF is expressed in the dorsal vagal complex (DVC) (box on the right), where it regulates energy metabolism as a downstream effector of the MC4R. BDNF is also expressed in the ventral tegmental area (VTA), where it is possibly involved in hedonic aspects of feeding (center box). *In contrast to the rest of the figure, this sentence refers to what happens when BDNF is deleted in the VTA. [Inset figure adapted from Paxinos and Watson (206).]

The PVN is a site for TRH synthesis and secretion, and TRH plays an important role in the control of energy homeostasis (149) through the TRH-TSH-tyrosine hydroxylase (TH) cascade. Triiodothyronine (T3) activates orexigenic neurons of the VMN and stimulates food intake independently of energy expenditure (138). TRH neurons express BDNF in response to immobilization stress (233). Interestingly, BDNF and T3 have opposing effects on the expression of several obesity-related genes in the hypothalamus (35). BDNF increases, while T3 decreases expression of BDNF, leptin receptor, proopiomelanocortin (POMC), TRH, and agouti-related protein (AgRP) (35). Through TrkB receptor signaling, BDNF increases the expression of the TRH precursor pre-pro-TRH mRNA in PVN neurons (260). Thus, in the PVN, BDNF may, in part, contribute to negative energy balance via increasing TRH, or, conversely, TRH may affect energy balance by increasing BDNF.

POMC and AgRP [neuropeptide Y (NPY/AgRP)] neurons are two distinct populations of neurons that project from the hypothalamic Arc to the PVN and release alpha-melanocyte stimulating hormone (α-MSH) and AgRP, respectively (71, 72), which interact with the melanocortin receptor (MC3/4R) in the PVN to elicit opposing actions on food intake. The MC4R is a membrane-bound α-MSH receptor (199), highly expressed in both the hypothalamus and brain stem (281), and is important for maintaining a lean phenotype. Animals heterozygous for, or lacking the MC4R gene become obese, (108), and MC4R agonist infusion reduces food intake and lowers body weight in HFD-fed rats (225). Delivery of BDNF gene into the hypothalamus increases the expression of MC4R in animals fed with regular chow or a HFD (39), suggesting BDNF affects melanocortin signaling (39). Notably, activation of MC4R leads to acute elevations of hypothalamic BDNF, which is critical for melanocortinergic effects on appetite and body temperature (196). BDNF is highly expressed in the VMN (262), which is an important area for regulating energy metabolism (220). MC4R controls BDNF expression in the VMN, and infusion of BDNF in the brain of MC4R-deficient mice attenuates hyperphagia and excessive weight gain induced by a moderate fat diet (25.1% calories from fat) (284). In addition, the phenotype of the TrkB homomorph, a mutant with reduced BDNF/TrkB signaling, is similar to the MC4R-null mutant mouse (284). The effects of BDNF in MC4R signaling are dependent on TrkB activation, as TrkB ligands reduce food intake and body weight downstream of MC4R in the hypothalamus of mice (255). Together, these data suggest that BDNF is a downstream effector of MC4R activation and that BDNF-TrkB signaling is an essential part of the mechanism for the anorectic and obesity-resistant effects of MC4R agonists in the VMN. (Fig. 1)

The VMN is involved in the control of autonomic responses that contribute to the prevention of obesity [reviewed by King (126)]. Neurons of the VMN project to many areas associated with feeding behavior, including the amygdala (37, 221), Arc (237), LH (224), PVN (157), the DMN (162, 245), as well as areas relating to rewarding aspects of feeding behavior, including the VTA (221), the NA, and the nucleus of the solitary tract (37). The VMN receives inputs from the Arc (15, 101), the LH (70, 221, 245), and the amygdala (161, 169). The VMN is involved in promoting satiety, as lesions in this area are associated with hyperphagic behavior (127). SF-1 is a nuclear hormone receptor important to the developmental structure of the VMN (109, 231). SF-1 is coexpressed with BDNF in the VMN and is involved in BDNF synthesis. Reduced levels of SF-1, as seen in heterozygous animals are associated with reduced BDNF, increased weight gain, hyperphagia, and lower daytime metabolic rate (254).

We found that a single injection of BDNF directly into the VMN, at doses not causing taste aversion, significantly decreases normal feeding and deprivation- and NPY-induced feeding for up to 48 h (269). No effects on feeding behavior were observed during the initial 4-h post injection, indicating that the feeding effects of injected BDNF might be indirect, or might take place as a result of retrograde (184, 185) or anterograde (7, 51, 234) transfer to a different brain location (269). The possibility that BDNF may act locally to reduce food intake by altering synaptic strength or receptor expression in the VMN has not been adequately investigated, but it is worth considering. In contrast to the delayed anorectic action, BDNF in the VMN immediately increased energy expenditure by elevating resting metabolic rate and physical activity (268). Chronic BDNF in the VMN also reduces HFD-induced obesity by reducing energy intake and/or increasing energy expenditure based on the phenotype of the animals on a HFD (272). Unlike in the PVN, BDNF in the VMN significantly increases physical activity in animals on regular chow (268) or HFD, suggesting that elevated physical activity-induced energy expenditure contributes to increased thermogenesis induced by BDNF in the VMN. BDNF in the VMN also decreases respiratory exchange ratio in animals on regular chow (268) and a HFD (272), indicating that BDNF stimulates fat metabolism, which partially explains the preferential loss of fat (vs. lean) mass after BDNF treatment. Deletion of BDNF in the VMN causes hyperphagia and obesity in mice (262), further confirming the importance of BDNF as a contributor to the maintenance of a lean phenotype. BDNF expression is responsive to dietary cues, as transcriptional levels of BDNF in the VMN are reduced by fasting (284) and elevated by glucose (262). It is possible that HFD-induced obesity results from lack of responsiveness to these dietary cues. In support of this, Yu et al. (292) observed decreased BDNF mRNA in the VMN of diet-induced obese mice, compared with mice resistant to HFD-induced obesity (292).

The adipokine leptin is produced and released from adipose tissue. Leptin signaling in the CNS inhibits food intake and increases energy expenditure, and in so doing, counters the accumulation of adiposity. Of note, leptin increases BDNF in certain brain areas, including the DVC (17), VMN, and DMN (137), with reductions in food intake (17, 18, 137, 268, 269) (Fig. 1). However, the antiobesity antidiabetic effects of BDNF on energy metabolism occur downstream of leptin signaling, as the effects are observed in leptin receptor deficient db/db (256–258), Kkay (an animal model of metabolic syndrome) (187), and diet-induced obese (DIO) mice (187, 255). Since leptin receptor signaling is important for leptin-dependent BDNF up-regulation (137), the obese phenotype might partially be related to the inability of leptin to increase hypothalamic BDNF expression. While leptin increases hypothalamic BDNF, BDNF decreases leptin production in adipocytes, an effect that involves sympathoneural β-adrenergic signaling and the hypothalamic-pituitary-adrenal axis (HPA) (40). HFD induces leptin resistance, characterized as a reduced anorectic response to leptin, as well as hyperleptinemia. We found that VMN BDNF significantly attenuated hyperleptinemia compared with that prior to BDNF intervention, or to vehicle-treated control animals (272) on a HFD. Additional studies are needed to explore whether the observed attenuated hyperleptinemia after BDNF is associated with or is an indication of improved leptin sensitivity.

Both BDNF and the TrkB receptor are highly expressed in the DVC of the hindbrain (51). The DVC is located in the caudal brain stem, an autonomic integrator of food intake control (17) and is involved in integrating satiety signals emanating from peripheral fat stores (148). BDNF acts as an anorexigenic factor in the DVC; Bariohay et al. (17) reported that BDNF infusion in the DVC induced anorexia and weight loss. Noteworthy, the efficacy of BDNF as an anorectic agent in the DVC decreased over a 14-day infusion period, indicating some compensation or desensitization occurs. The protein content of BDNF in the DVC decreases after 48 h of food deprivation and increases upon refeeding. Furthermore, the anorexigenic hormones leptin and CCK injected peripherally increase BDNF content in the DVC (17). The DVC contains the neural network responsible for the central pattern generator of swallowing (111) and BDNF-TrkB signaling inhibits the swallowing reflex via modulation of GABAergic signaling (19). Similarly, increased stimulation of the superior laryngeal nerve decreases BDNF in the DVC, indicating positive feedback allowing the swallowing reflex to continue with the presence of a food stimulus (19, 148). BDNF is a downstream effector of the MC4R signaling pathway in the DVC and is necessary for the anorexigenic effect of MC4R activation (18). The orexigenic effect of an MC4R antagonist is abolished with coadministration of BDNF, and pharmacological blockade of the TrkB receptor attenuates the anorexigenic effect of an MC4R agonist (18). Taken together, BDNF in the DVC appears to be responsive to hormonal satiety signals, as well as physical signals of the presence of a food stimulus, and coordinates swallowing. (Fig. 1)

MC4R is expressed in the amygdala (181) an area involved in regulating macronutrient selection (128) and some reward aspects of feeding behavior (75, 122). In the amygdala, injection of MC4R agonist causes a dose-dependent reduction in food intake, which is greater in animals fed a HFD (32). Surprisingly, while injections of the orexigenic hormone AgRP in the amygdala increases food intake, it is also associated with elevated amygdala BDNF mRNA (32). This effect is unexpected and warrants further investigation, as compelling evidence suggests that BDNF is a downstream effector of anorexigenic melanocortinergic signaling (32).

BDNF and TrkB are expressed in the mesolimbic dopamine system, which is associated with hedonic reward (198, 229). The consumption of palatable high-fat foods alters the expression of BDNF and TrkB receptor in the VTA, but not the nucleus accumbens (NAc) of wild-type mice (52). BDNF is not highly expressed in the NAc. Most of the BDNF found in the NAc is produced in the VTA and anterogradely transported from neurons that originate there (51, 198). The neurons of the NAc release dopamine in response to palatable foods (22). Site-specific viral depletion of BDNF in the VTA causes excessive intake of a palatable HFD, but not standard chow, whereas reduced BDNF in the VMN results in indiscriminate hyperphagia of either HFD or chow (52, 262). Peripheral administration of a D1 receptor agonist normalizes the caloric intake of palatable HFD in BDNF mutant mice (52), indicating that BDNF synthesis in the VTA is possibly involved in dopamine secretion from neurons of the NAc and thus BDNF may play a role in hedonic reward. (Fig. 1)

The hippocampus, which has long been associated with learning and memory (110, 236), has been implicated in having a potential involvement in energy balance (58). Part of what makes this an attractive hypothesis is that areas of the hippocampus, in particular, field CA1 neurons, project to the LH, Arc, PVN, DMN, and VMN, all important hypothalamic areas involved in feeding behavior (44). Several multisynaptic pathways have been identified that connect brain stem feeding control areas to the hippocampus (96, 180). Amnesic patients with hippocampus damage showed reduced sensitivity to interoceptive signals of hunger and satiety (102, 218). Compared with intact controls, rodents with selective lesions of the hippocampus exhibit increased appetitive responding for food (49, 59, 223). In a recent study by Davidson et al. (57), lesioning of the complete hippocampus resulted in increased food intake, body weight gain, appetitive behavior, and metabolic activity. When lesioning was restricted to the ventral pole, which projects to the lateral hypothalamus (44), animals had increased food intake and body weight (57). Functional magnetic resonance imaging (fMRI) identified the hippocampus and prefrontal cortex as the sites of greatest activation in obese people (273). DelParigi et al. (60) also noticed a decreased hippocampal blood flow in obese and formerly obese people after they consumed a liquid meal to satiation. These findings suggest that the hippocampus plays an important role in the regulation of energy metabolism. BDNF and TrkB are highly expressed in the hippocampus. Many studies have reported that exercise increases hippocampal BDNF expression (47, 95, 227) and that these increases are associated with enhanced cognition (26, 143, 195). Exactly what role, if any, hippocampal BDNF plays in energy metabolism is still unclear. Some evidence suggests that hippocampal BDNF might be related to factors affecting the memory of food, and, therefore, motivation to eat (84). A/J mice, who behaviorally model activity-induced anorexia and exhibit reduced food anticipatory activity, have significantly lower BDNF expression in the hippocampus during feeding times, compared with mice that have normal food anticipatory activity (84). Dietary restriction has no effect on hippocampal BDNF in A/J mice, whereas in mice without activity-induced anorexia, dietary restriction increases hippocampal BDNF (62, 84, 150), as well as increasing the full-length TrkB receptor (150). This is likely not directly due to altered glucose levels, as no changes in BDNF expression were observed in the hippocampus with intracerebroventricular glucose administration (262). A HFD decreases hippocampal BDNF (178, 204, 278280), as does a HF-high sugar diet (118, 177, 239); however, the type of sugar matters as the combination of a HFD and dextrose decreased BDNF, whereas HFD and sucrose did not. In this study, rats gained similar amounts of weight, indicating the differences in BDNF expression were not directly related to body weight (118). Furthermore, a HFD does not always decrease hippocampal BDNF. In an animal model of early life trauma, in which rats were separated from dams for about 2 wk after birth, a HFD was associated with increased hippocampal BDNF (166). Yu et al. (292) observed that in animals prone to DIO, hippocampal BDNF was reduced in response to a HFD, whereas in obesity-resistant animals (DRO), or in pair-fed DIO animals, it was not. The authors speculate that the decreases in hippocampal BDNF signaling may correspond to weakened inhibitory control of HF food intake and promote obesity (292). Davidson et al. (58) proposed a “vicious circle” model: an unhealthy diet (such as HFD) reduces hippocampal BDNF, which causes hippocampal dysfunction (such as hypermnesia), which may result in impaired feeding behavior and overeating in an “obesigenic” environment, which further reduces hippocampal BDNF and damages hippocampal function, and continues the circle (58).

Peripheral Actions of BDNF

In addition to the brain, BDNF is expressed in many tissues important to the regulation of energy homeostasis, namely adipose tissue, skeletal and smooth muscle, and liver (42, 159, 182, 261). It is, therefore, important to consider that the effect of BDNF in these peripheral tissues might also contribute to the overall maintenance of energy balance.

BDNF mRNA and protein expression are increased in exercising skeletal muscle, an effect associated with the phosphorylation of AMPK and acetyl-coA carboxylase β (ACCβ), as well as increases in fatty acid oxidation (170). AMPK “senses” a high-energy state in muscle, and when activated, it phosphorylates the mitochondrial ACCβ, thereby inhibiting increases in malonyl-CoA levels (an action that ultimately leads to increased mitochondrial fatty acid transport and oxidation) (41, 172, 173). BDNF elevates malonyl-CoA independently of AMPK; however, the effect of BDNF on muscle cell fatty acid oxidation is AMPK dependent (170). Skeletal muscle has a high nutritive demand, and poor intramuscular fatty acid metabolism may contribute to obesity (106) and insulin resistance (107, 222). The finding that BDNF increases ACCβ phosphorylation and fatty acid oxidation may be one of the ways in which peripheral BDNF increases insulin sensitivity and weight loss. Muscle-derived BDNF is not released into circulation (170), and although exercise is known to increase circulating levels of BDNF (74), the source of this increase is unlikely from muscle cells (170).

In the liver, BDNF contributes to the development of hyperglycemia, hyperinsulinemia, elevated serum cholesterol, and triglycerides associated with eating a HFD (242). When fed a HFD, liver-specific BDNF-knockout mice have elevated levels of peroxisome proliferator-activated receptor alpha (PPARα) and fibroblast growth factor 21 (Fgf21) compared with wild-type mice (242). Fgf21 is a downstream target of PPARα, which is important for hepatic lipid oxidation, and insulin sensitivity (14, 125, 285). Fgf21 dose dependently reduces body weight and adiposity and reduces the expression of a variety of genes involved in fatty acid and triglyceride synthesis (285). The protection against HFD-induced hyperglycemia and hyperinsulinemia observed in liver-specific knockouts is inconsistent with previous research, which reported improvements in liver histology after subcutaneous BDNF treatment (288). Thus, the improved liver histology viewed in these cases is likely secondary to other peripheral effects of BDNF.

Mature BDNF (~13.6 kDa), which is less than half the size of pro-BDNF (~32 kDa), does not cross the blood-brain barrier (203), and therefore, studies reporting physiological effects of subcutaneous injections of BDNF may be reflective of peripheral actions. Subcutaneous BDNF enhances glucose utilization in muscle and BAT of db/db mice, but not normoglycemic animals (290). Additionally, BDNF restores levels of insulin-secreting granules in beta cells and maintains their histologic cellular organization in db/db mice, even though there is no TrkB receptor in pancreatic islets (286). Subcutaneous injections of BDNF reduced body weight, fat pad weight, and liver weight in db/db mice compared with pair-fed Troglitazone-treated db/db mice (288). A single injection of BDNF has a prolonged hypoglycemic effect, which is not attributable to reductions in food intake alone (289). Taken together, this indicates that BDNF may act peripherally to normalize blood glucose in hyperglycemic rodents, or in those no longer sensitive to leptin, possibly by enhancing muscle utilization or via the stimulation of BAT.

While both central and peripheral activation of the TrkB receptor reduces food intake and obesity in rodents, this effect is not conserved across all species. In monkeys, the response of peripheral TrkB activation is orexigenic and obesity promoting, while the central activation parallels the anorectic effects observed in mice (156). The involvement of BDNF in energy homeostasis in humans is difficult to study, and studies are usually limited to correlations between serum or plasma BDNF and body weight or adiposity. Serum BDNF levels are lower in human type-2 diabetic patients (82), AN (191, 219), and in extremely overweight children (66), but not in bulimia nervosa (219), recovered AN (191), or healthy controls (191, 219). Thus, the relationship between serum BDNF and BMI is not clear. In one study, there was a significant positive relationship between the two (219), and in another, serum BDNF negatively correlated with both BMI and body fat (66). It is likely that serum BDNF reflects the amount stored in platelets, which is released during the clotting process (81, 174). This level is not acutely altered by food intake (66). Plasma BDNF, on the other hand, likely has many sources (34, 85, 124, 189). Plasma BDNF is higher in obese women, but these levels are significantly dropped after bariatric surgery (175). Conversely, Mercader et al. (174) observed no correlation between plasma BDNF and BMI in clinical subgroups of eating disorder patients. Krabbe et al. (141) report that plasma levels of BDNF are decreased in human type 2 diabetics, independent of obesity. By sampling from the internal jugular vein and comparing plasma BDNF from arterial and venous samples, Krabbe et al. (141) demonstrated that cerebral output of BDNF is reflected in the circulation. They observed that plasma BDNF levels are directly, inversely related to fasting plasma glucose levels and that BDNF output from the brain is inhibited when blood glucose levels are elevated.

BDNF and Neuronal Plasticity

Neuronal plasticity is defined as an experience-dependent change in synaptic strength (31). A well-studied example of this is long-term potentiation (LTP), which has particularly been associated with neurons of the hippocampus. LTP is the activity-dependent strengthening of a synapse, which is typically induced by high-frequency stimulation of excitatory input. Many studies have identified the importance of BDNF in the development of LTP (76, 117, 139, 205, 283, 293) and also in the development from LTP to long-term memories (6, 249, 250). BDNF is secreted in response to a high-frequency stimulation and is dependent on Ca2+ influx through voltage-gated Ca2+ channels or NMDA receptors (2, 16, 100). BDNF can bind TrkB receptors on either side of the synapse (61). In the hippocampus, it functions to facilitate the presynaptic release of excitatory neurotransmitters (33), as well as postsynaptic AMPA receptor insertion (154, 155) and dendritic spine maintenance (45, 55, 240). In contrast to LTP promotion by mature BDNF, recent evidence indicates a potential role for pro-BDNF in facilitating long-term depression (LTD) through activation of p75NTR receptor (69, 160, 277).

Factors related to energy balance have been described to affect hippocampal LTP. When fed a diet high in both fat and sucrose (HFS), rats were less capable in a spatial learning capacity task than rats fed standard chow (278). There was no sign of neuronal degeneration in the HFS fed rats. Additionally, rats with the lowest hippocampal BDNF, who also had lower levels of CREB and the vesicle-associated synapsin I (114), had the lowest learning capacity (278). A HFD also has a negative impact on hippocampal LTP and plasticity (210), and some studies suggest that this impairment is related to BDNF. A diet high in saturated fat and refined sugars significantly reduced hippocampal BDNF compared with rats fed standard chow, which was accompanied by impaired LTP and poor performance on the Morris water maze, a spatial learning task (177). A diet high in saturated fat reduced BDNF in both the ventral hippocampus and medial prefrontal cortex with associated impairment in reversal learning (118). Taken together, these data suggest that a HFD may affect synaptic plasticity in rat hippocampal neurons. Conversely, caloric restriction (77) and exercise (238) enhance LTP and are associated with increased BDNF (150, 194). Hippocampal BDNF levels are increased with running (53), and running protects against stress-related downregulation of BDNF (1).

Although the role of BDNF in neuronal plasticity has been well studied in the hippocampus, the possibility that BDNF contributes to plasticity of hypothalamic neurons related to energy balance is less well studied. The molecular mechanisms by which BDNF acts in the hypothalamus to affect energy homeostasis have not been characterized. However, an example of BDNF-dependent hypothalamic plasticity has been recently described in thermal sensation and temperature control development (120). Several plasticity-related genes (including BDNF) were reported to be differentially expressed in high and low body weight chickens during development, suggesting different inherent capacities of these animals to adapt appetite circuitry (115). We have recently observed reductions in feeding behavior in rats after BDNF administration into the VMN and PVN. We observed this feeding inhibition between 4 and 24 and 24 and 48 h postinjection in both sites; however, we did not observe an effect of BDNF during the first 4 h (267, 269). The possibility that BDNF acts to suppress food intake via a plasticity-related mechanism in the hypothalamus is thus worth investigating. Neuronal activity in the PVN is modulated by excitatory glutamatergic signaling, as well as other excitatory and inhibitory neurotransmitters. Recently, NMDA receptor-dependent plasticity in the PVN has been described to reduce excitatory glutamatergic signaling in spontaneously hypertensive rats (153), and the NMDA receptor-dependent plasticity was induced environmentally by chronic intermittent hypoxia (50). We have observed that a single injection of BDNF in the PVN dose dependently suppresses feeding for up to 48 h (267). The duration of the effect of a single injection of BDNF into the PVN, as well as characterization of plasticity mechanisms in the PVN, makes neural plasticity seem a candidate mechanism for the observed anorectic response.

BDNF and Neurogenesis

The adult central nervous system contains neuronal stem cells capable of generating new neurons (8), and several progenitor cells in certain brain regions have been identified (89, 243). Neurogenesis has been well studied in the subventricular zone of the lateral ventricles, and the subgranular zone of the hippocampal formation (83). In the hippocampus, Nakatomi et al. (190) observed the regeneration of CA1 pyramidal neurons of the adult rodent brain following ischemic degeneration, which was facilitated by the intracerebroventricular infusion of two different growth factors (FGF-2 and EGF). Novel to this study was the observation that growth factors signal the recruitment of progenitors from areas near the hippocampus to facilitate neurogenesis in areas where no progenitors are available. Thus, extensive neurogenesis is possible in brain sites not containing stem cells (190). Mitosis of progenitor cells lasts about 24 h for the rat (36), and 14 h in the mouse (165), after which it takes 3–4 wk for new neurons to mature and fully integrate into the circuitry (263). Thus, any effect of BDNF on neurogenesis would likely be observable more over the long term.

Energy balance has been described to affect BDNF and neurogenesis in the hippocampus. Dietary restriction (DR) increases neurogenesis in the adult mouse hippocampus, an effect associated with elevated BDNF expression, yet this effect is absent in BDNF heterozygous mice (150). However, DR normalizes BDNF in the hippocampus, striatum, and cerebral cortex of Huntington mutant mice, in whom BDNF expression is decreased, such that it is equivalent to the expression in wild-type ad libitum-fed mice, and reverses obesity, abnormal locomotor activity, and hyperphagia (62). Conversely, a diet high in saturated fat decreases BDNF and compromises cognitive performance (177). In mice susceptible to DIO, a HFD decreases BDNF and TrkB mRNA in the hippocampus compared with mice resistant to DIO (292). Exposure to enriched environment and exercise elevates BDNF in the hippocampus and improves learning (200). Conversely, it has been observed in rats that the consumption of a HFD, particularly one high in saturated fat, decreases BDNF and adult hippocampal neurogenesis (204). Park et al. (204) observed that the prevention of neurogenesis was associated with increased levels of malondialdehyde (MDA) in the hippocampus, which is an indicator of lipid peroxidation, and they found a direct effect of MDA administration on inhibiting neurogenesis. A HFD fed to dams is associated with obesity and hyperlipidemia in offspring. These offspring have impaired hippocampal neurogenesis, which parallels with the degree of neuronal impairment observed when treating cells in vitro with MDA (253). Tozuka et al. (252) observed that a maternal HF diet caused reductions in hippocampal BDNF and impaired dendritic arborization of hippocampal neurons in pups.

Neurogenesis in the hypothalamus is less well characterized. The effect of a maternal HFD on hypothalamic neurogenesis in pups has recently been investigated, where HFD increased neurogenesis of neurons expressing orexigenic peptides galanin, enkephalin, and dynorphin in the PVN and orexin and melanin-concentrating hormone in the perifornical lateral hypothalamus (46). These hypothalamic changes were associated with increased body weight, leptin, insulin, dietary fat preference, triglycerides, and galanin expression in the PVN (46). Thus, it appears that energy balance affects neurogenesis; however, a compelling question that remains to be addressed is whether neurogenesis also affects energy balance.

Neuronal stem cells have been identified in the hypothalamus, and hypothalamic neurogenesis has been described to occur at a low rate (134). Ciliary neurotrophic factor (CNTF) is similar to BDNF in that it promotes neuronal survival (146, 228) and the maintenance of neuronal stem cells (230), and it activates signaling cascades in the hypothalamus involved in feeding and energy homeostasis (30, 144). Kokoeva et al. (135) observed an increase in neurogenesis in several hypothalamic feeding centers, particularly the median eminence-Arc, in CNTF-treated mice fed a HFD compared with non-CNTF-treated controls. Furthermore, the CNTF-treated mice were resistant to weight gain on the HFD, which persisted for more than a week after the cessation of CNTF treatment. The antimitotic agent cytosine-B-d-arabinofuranoside prevented the inhibition of weight gain after the cessation of CNTF treatment, but not during, indicating neurogenesis is partially responsible for the sustained antiobesigenic effect of CNTF (135). Additionally, the new neurons expressed proteins involved in energy balance such as POMC, neuropeptide Y, and phosphorylated STAT3, an indicator of leptin signaling (135). That CNTF is capable of inducing hypothalamic neurogenesis and that this neurogenesis has an antiobesigenic effect is a demonstration that hypothalamic neurogenesis can, in fact, play an important role in regulating energy metabolism.

Using BrdU, Pencea et al. (208) were the first to observe that BDNF is capable of inducing hypothalamic neurogenesis after continuous intracerebroventricular administration of BDNF for 12 days. Neurogenesis triggered by BDNF correlates with levels of TrkB expression, but the TrkB receptor is not directly integrated into new neurons. Thus, hypothalamic regions that contained high levels of the TrkB receptor (e.g., the PVN) had greater amounts of BrdU cells than areas with less TrkB expression, even if those areas were located closer to the site of BDNF infusion (208). Recently, Kumar et al. (142) used BrdU to measure neurogenesis in response to DR during a model of cytotoxic injury. Alternate-day DR was associated with increased BDNF and neurogenesis in several brain regions, including the median eminence-Arc of the hypothalamus (142). The type of new neurons generated in the Arc with dietary restriction remains to be elucidated, as well as precisely how they might affect energy balance.

Neuroprotection and Survival

During development, neurotrophic factors are critical to survival because they inhibit apoptosis of developing neurons (282). BDNF is neuroprotective in the hippocampus, particularly against ischemic damage (23, 133, 147). Neuronal apoptosis, or programmed cell death, is characterized by the activation of caspases, specifically caspase 9, which acts upstream of caspase 3 (246). BDNF reduces glutamate-induced apoptotic cell death upstream of the activation of caspase-3-like enzymes and increases expression of the antiapoptotic protein B-cell lymphoma 2 (Bcl-2) (5). Additional studies have shown BDNF induces increases in Bcl-2 (12, 129, 209). It is important to note, however, that BDNF is not always protective. To some cultured neurons of the hippocampus and cerebrocortex BDNF is toxic (78). While activation of TrkB receptors increases LTP and neuronal survival, activation of p75NTR can lead to apoptosis (20) and LTD (277).

Oxidative stress (112) often leads to a loss of cell function, apoptosis, or necrosis [reviewed by Azad (13)]. BDNF has been implicated in being protective against oxidative stress by preventing the accumulation of peroxides and increasing antioxidant enzymes in hippocampal neurons (171). An abundant oxidative stress marker is 4-hydroxynonenal (HNE), which is generated through peroxidation of omega 6-polyunsaturated fatty acids (25, 68). HNE is highly diffusible and may contribute to oxidative stress far from a site of injury (68). Local application of BDNF on the dorsal hemisected spinal cord in rats resulted in reduced lipid peroxidation, as shown by decreased HNE-immunoreactive staining (113). This effect was observed within 48 h of BDNF application and was associated with a reduction in activated microglial cells, which may have contributed to decreased oxidative stress (113).

In brain regions exhibiting neuronal loss, as is associated with Huntington's disease (73) and Alzheimer's disease (104), BDNF levels are low, which contributes to the neuronal degeneration associated with these diseases (104). Tg2576 mice are a widely used model of Alzheimer's disease. They develop amyloid plaques and have declining cognitive function at 6 mo old (274). Kohjima et al. (132) observed Tg2576 mice fed a HFD developed obesity and insulin resistance due to hyperphagia, and that the abnormal feeding behavior was associated with increased amyloid plaque formation and decreased hypothalamic BDNF. Oxidative stress is present early in the pathogenesis of Alzheimer's disease (121, 215). Vitamin E is a known antioxidant, which can prevent the development of oxidative stress. Vitamin E supplementation ameliorates HFD-induced reductions in BDNF, suggesting that BDNF levels may be altered in response to oxidative stress (279). Several studies have identified that there is a relationship between a HFD and impairments in cognitive performance, particularly in rats fed a diet high in saturated fats (92, 93, 275, 276). In addition to affecting cognitive function, a HFD induces apoptosis of hypothalamic neurons, such as POMC neurons (179). Thus, the neuroprotective effect of BDNF may have a role in the central regulation of energy metabolism; however, further studies are needed to define the role of BDNF in hypothalamic neuroprotection and how it relates to energy balance.

Factors Affecting Expression of Hypothalamic BDNF

Hypothalamic BDNF and TrkB content are affected by age. In rats raised in standard laboratory conditions, the mature form of hypothalamic BDNF peaks at about 1 wk postnatally in rats, remains elevated for the first month of life, and declines with age (232). Additionally, declining levels of TrkB receptor begin in rats at around 2 mo, with extreme reductions observed by 22 mo (232). It is possible that environmental factors could prevent age-related decline in BDNF. Recently, Cao et al. (40) reported BDNF expression is increased in mice housed in an enriched environment. The environmental enrichment included a large open space, toys, and a running wheel. Earlier studies have shown that in this type of complex environment, animals exhibit improved learning and memory and increased neurogenesis (38, 63). Additionally, mice living in an enriched environment remain leaner than mice in standard housing, an effect associated with consistently elevated BDNF expression in the Arc (40). Elevated BDNF in the hypothalamus associated with environmental enrichment decreased leptin levels, which inhibited cancer tumor growth in several models of cancer. The metabolic profiles of mice in an enriched environment were mimicked when BDNF was overexpressed in the hypothalamus of animals living in standard housing using a recombinant adeno-associated virus vector (40).

Four weeks of running increased slightly, but not significantly, expression of BDNF in the Arc, whereas an enriched environment significantly increased Arc BDNF expression after only 2 wk. While both the running mice and the enriched environment mice had similar decreases in body weight, the groups differed in metabolic gene expression, and only the enriched environment group had a corresponding decrease in leptin and tumor growth (40). The study by Cao et al. (40) opens the door to many questions about the relationship between environment and metabolism, more specifically, how does environmental enrichment impact food intake and energy expenditure to contribute to obesity resistance? What constitutes an enriched environment? In another model of environmental enrichment, Angelucci et al. (9) report that music increased BDNF in the mouse hypothalamus.

Acute immobilization stress causes rapid increases in hypothalamic BDNF (186, 213). These increases are accompanied by decreased body weight and increased locomotor activity, as well as activation of the HPA (186). Conversely, neonatal stress induced by separating rat pups from their dams for 180 min per day is associated with elevations in hippocampal BDNF protein expression, which is likely exerting a protective effect on existing neurons there (94).

Kohjima et al. (132) report that 16 wk of a HFD, which was sufficient to induce insulin resistance, obesity, and amyloid plaques in Tg2567 mice, decreased hypothalamic BDNF and led to elevated feeding behavior. In the VMN specifically, BDNF, but not TrkB mRNA, is lower in DIO mice compared with DRO, suggesting that the DIO phenotype may, in part, be mediated through low BDNF expression in this important feeding center (292). No differences in BDNF expression were observed between the two phenotypes in the Arc, the DMN or the PVN (292). However, Sprague-Dawley rats fed either a high-energy diet (HE) or high energy plus Ensure liquid diet (HE + E) expressed increased TrkB receptor, but not BDNF, in the VMN (10). Both groups decreased food intake and body weight when switched from their respective diets to standard chow, and in both cases, this was accompanied by decreased BDNF. The HE, but not the HE+E group also had decreased TrkB expression upon switching to standard chow (10). Differences in TrkB mRNA are difficult to interpret, because not all forms of TrkB are active and some might serve to inhibit TrkB signaling. We have recently demonstrated that chronic administration of BDNF into the PVN reduced HFD-induced obesity and that animals with greater body fat were more responsive to the effect of added BDNF (270). This is in line with recent evidence that suggests that animals resistant to HFD-induced obesity maintain higher basal levels of BDNF and TrkB in the hypothalamic VMN compared with animals susceptible to DIO (292). Moreover, VMN BDNF levels were further reduced in DIO mice on a HF diet, and they were negatively correlated with plasma glucose and positively correlated with plasma adiponectin (292). Adiponectin is an adipokine associated with increased insulin sensitivity and reduced appetite (27). This suggests that BDNF and TrkB expression in the hypothalamus might play a role in determining susceptibility to obesity (292). The antiobesity effect of BDNF in the PVN was related to a significant reduction in energy intake (270), and previously, we have observed that BDNF in the PVN increases energy expenditure and resting metabolic rate (266).

Perspectives and Significance

BDNF deficiency is associated with increased weight in mice and humans, and BDNF administration in the hypothalamus can reduce food intake and increase energy expenditure, leading to lighter animals. The two critical hypothalamic sites are the PVN and VMN, but other brain and peripheral sites may also play a role. There is strong evidence that BDNF in the hippocampus is involved in neural plasticity and neurogenesis in adult animals, but whether hypothalamic BDNF exerts its energy balance effects through plasticity and neurogenesis has not yet been determined. There is evidence that the BDNF-induced negative energy balance persists long after BDNF administration, which would be compatible with plastic mechanisms. Much more investigation is needed concerning whether hypothalamic BDNF mediates energy balance, in part, via plasticity and/or neurogenesis, and whether hippocampal and hypothalamic BDNF are influenced by environment and energy states.

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the author(s).

REFERENCES

1. Adlard PA, Cotman CW. Voluntary exercise protects against stress-induced decreases in brain-derived neurotrophic factor protein expression. Neuroscience 124: 985–992, 2004 [Abstract] [Google Scholar]
2. Aicardi G, Argilli E, Cappello S, Santi S, Riccio M, Thoenen H, Canossa M. Induction of long-term potentiation and depression is reflected by corresponding changes in secretion of endogenous brain-derived neurotrophic factor. Proc Natl Acad Sci USA 101: 15788–15792, 2004 [Europe PMC free article] [Abstract] [Google Scholar]
3. Alderson RF, Alterman AL, Barde YA, Lindsay RM. Brain-derived neurotrophic factor increases survival and differentiated functions of rat septal cholinergic neurons in culture. Neuron 5: 297–306, 1990 [Abstract] [Google Scholar]
4. Allendoerfer KL, Cabelli RJ, Escandon E, Kaplan DR, Nikolics K, Shatz CJ. Regulation of neurotrophin receptors during the maturation of the mammalian visual system. J Neurosci 14: 1795–1811, 1994 [Abstract] [Google Scholar]
5. Almeida RD, Manadas BJ, Melo CV, Gomes JR, Mendes CS, Graos MM, Carvalho RF, Carvalho AP, Duarte CB. Neuroprotection by BDNF against glutamate-induced apoptotic cell death is mediated by ERK and PI3-kinase pathways. Cell Death Differ 12: 1329–1343, 2005 [Abstract] [Google Scholar]
6. Alonso M, Vianna MR, Izquierdo I, Medina JH. Signaling mechanisms mediating BDNF modulation of memory formation in vivo in the hippocampus. Cell Mol Neurobiol 22: 663–674, 2002 [Abstract] [Google Scholar]
7. Altar CA, Cai N, Bliven T, Juhasz M, Conner JM, Acheson AL, Lindsay RM, Wiegand SJ. Anterograde transport of brain-derived neurotrophic factor and its role in the brain. Nature 389: 856–860, 1997 [Abstract] [Google Scholar]
8. Altman J. Are new neurons formed in the brains of adult mammals? Science 135: 1127–1128, 1962 [Abstract] [Google Scholar]
9. Angelucci F, Ricci E, Padua L, Sabino A, Tonali PA. Music exposure differentially alters the levels of brain-derived neurotrophic factor and nerve growth factor in the mouse hypothalamus. Neurosci Lett 429: 152–155, 2007 [Abstract] [Google Scholar]
10. Archer ZA, Rayner DV, Barrett P, Balik A, Duncan JS, Moar KM, Mercer JG. Hypothalamic energy balance gene responses in the Sprague-Dawley rat to supplementation of high-energy diet with liquid ensure and subsequent transfer to chow. J Neuroendocrinol 17: 711–719, 2005 [Abstract] [Google Scholar]
11. Arija V, Ferrer-Barcala M, Aranda N, Canals J. BDNF Val66Met polymorphism, energy intake and BMI: a follow-up study in schoolchildren at risk of eating disorders [Online]. BMC Public Health 10: 363, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
12. Assuncao M, Santos-Marques MJ, Carvalho F, Andrade JP. Green tea averts age-dependent decline of hippocampal signaling systems related to antioxidant defenses and survival. Free Radic Biol Med 48: 831–838, 2010 [Abstract] [Google Scholar]
13. Azad N, Iyer A, Vallyathan V, Wang L, Castranova V, Stehlik C, Rojanasakul Y. Role of oxidative/nitrosative stress-mediated Bcl-2 regulation in apoptosis and malignant transformation. Ann NY Acad Sci 1203: 1–6, 2010 [Abstract] [Google Scholar]
14. Badman MK, Pissios P, Kennedy AR, Koukos G, Flier JS, Maratos-Flier E. Hepatic fibroblast growth factor 21 is regulated by PPARα and is a key mediator of hepatic lipid metabolism in ketotic states. Cell Metab 5: 426–437, 2007 [Abstract] [Google Scholar]
15. Bagnol D, Lu XY, Kaelin CB, Day HE, Ollmann M, Gantz I, Akil H, Barsh GS, Watson SJ. Anatomy of an endogenous antagonist: relationship between Agouti-related protein and proopiomelanocortin in brain [Online]. J Neurosci 19: RC26, 1999 [Abstract] [Google Scholar]
16. Balkowiec A, Katz DM. Cellular mechanisms regulating activity-dependent release of native brain-derived neurotrophic factor from hippocampal neurons. J Neurosci 22: 10399–10407, 2002 [Abstract] [Google Scholar]
17. Bariohay B, Lebrun B, Moyse E, Jean A. Brain-derived neurotrophic factor plays a role as an anorexigenic factor in the dorsal vagal complex. Endocrinology 146: 5612–5620, 2005 [Abstract] [Google Scholar]
18. Bariohay B, Roux J, Tardivel C, Trouslard J, Jean A, Lebrun B. Brain-derived neurotrophic factor/tropomyosin-related kinase receptor type B signaling is a downstream effector of the brainstem melanocortin system in food intake control. Endocrinology 150: 2646–2653, 2009 [Abstract] [Google Scholar]
19. Bariohay B, Tardivel C, Pio J, Jean A, Felix B. BDNF-TrkB signaling interacts with the GABAergic system to inhibit rhythmic swallowing in the rat. Am J Physiol Regul Integr Comp Physiol 295: R1050–R1059, 2008 [Abstract] [Google Scholar]
20. Barrett GL. The p75 neurotrophin receptor and neuronal apoptosis. Prog Neurobiol 61: 205–229, 2000 [Abstract] [Google Scholar]
21. Bartanusz V, Jezova D, Bertini LT, Tilders FJ, Aubry JM, Kiss JZ. Stress-induced increase in vasopressin and corticotropin-releasing factor expression in hypophysiotrophic paraventricular neurons. Endocrinology 132: 895–902, 1993 [Abstract] [Google Scholar]
22. Bassareo V, Di Chiara G. Differential influence of associative and nonassociative learning mechanisms on the responsiveness of prefrontal and accumbal dopamine transmission to food stimuli in rats fed ad libitum. J Neurosci 17: 851–861, 1997 [Abstract] [Google Scholar]
23. Beck T, Lindholm D, Castren E, Wree A. Brain-derived neurotrophic factor protects against ischemic cell damage in rat hippocampus. J Cereb Blood Flow Metab 14: 689–692, 1994 [Abstract] [Google Scholar]
24. Beckers S, Peeters A, Zegers D, Mertens I, Van Gaal L, Van Hul W. Association of the BDNF Val66Met variation with obesity in women. Mol Genet Metab 95: 110–112, 2008 [Abstract] [Google Scholar]
25. Benedetti A, Comporti M, Esterbauer H. Identification of 4-hydroxynonenal as a cytotoxic product originating from the peroxidation of liver microsomal lipids. Biochim Biophys Acta 620: 281–296, 1980 [Abstract] [Google Scholar]
26. Berchtold NC, Castello N, Cotman CW. Exercise and time-dependent benefits to learning and memory. Neuroscience 167: 588–597, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
27. Berg AH, Combs TP, Du X, Brownlee M, Scherer PE. The adipocyte-secreted protein Acrp30 enhances hepatic insulin action. Nat Med 7: 947–953, 2001 [Abstract] [Google Scholar]
28. Berkemeier LR, Winslow JW, Kaplan DR, Nikolics K, Goeddel DV, Rosenthal A. Neurotrophin-5: a novel neurotrophic factor that activates trk and trkB. Neuron 7: 857–866, 1991 [Abstract] [Google Scholar]
29. Bishop JF, Mueller GP, Mouradian MM. Alternate 5′ exons in the rat brain-derived neurotrophic factor gene: differential patterns of expression across brain regions. Brain Res Mol Brain Res 26: 225–232, 1994 [Abstract] [Google Scholar]
30. Bjorbaek C, Elmquist JK, El-Haschimi K, Kelly J, Ahima RS, Hileman S, Flier JS. Activation of SOCS-3 messenger ribonucleic acid in the hypothalamus by ciliary neurotrophic factor. Endocrinology 140: 2035–2043, 1999 [Abstract] [Google Scholar]
31. Bliss TV, Collingridge GL. A synaptic model of memory: long-term potentiation in the hippocampus. Nature 361: 31–39, 1993 [Abstract] [Google Scholar]
32. Boghossian S, Park M, York DA. Melanocortin activity in the amygdala controls appetite for dietary fat. Am J Physiol Regul Integr Comp Physiol 298: R385–R393, 2010 [Abstract] [Google Scholar]
33. Boulanger LM, Poo MM. Presynaptic depolarization facilitates neurotrophin-induced synaptic potentiation. Nat Neurosci 2: 346–351, 1999 [Abstract] [Google Scholar]
34. Braun A, Lommatzsch M, Mannsfeldt A, Neuhaus-Steinmetz U, Fischer A, Schnoy N, Lewin GR, Renz H. Cellular sources of enhanced brain-derived neurotrophic factor production in a mouse model of allergic inflammation. Am J Respir Cell Mol Biol 21: 537–546, 1999 [Abstract] [Google Scholar]
35. Byerly MS, Simon J, Lebihan-Duval E, Duclos MJ, Cogburn LA, Porter TE. Effects of BDNF, T3, and corticosterone on expression of the hypothalamic obesity gene network in vivo and in vitro. Am J Physiol Regul Integr Comp Physiol 296: R1180–R1189, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
36. Cameron HA, McKay RD. Adult neurogenesis produces a large pool of new granule cells in the dentate gyrus. J Comp Neurol 435: 406–417, 2001 [Abstract] [Google Scholar]
37. Canteras NS, Simerly RB, Swanson LW. Organization of projections from the ventromedial nucleus of the hypothalamus: a Phaseolus vulgaris-leucoagglutinin study in the rat. J Comp Neurol 348: 41–79, 1994 [Abstract] [Google Scholar]
38. Cao L, Jiao X, Zuzga DS, Liu Y, Fong DM, Young D, During MJ. VEGF links hippocampal activity with neurogenesis, learning and memory. Nat Genet 36: 827–835, 2004 [Abstract] [Google Scholar]
39. Cao L, Lin EJ, Cahill MC, Wang C, Liu X, During MJ. Molecular therapy of obesity and diabetes by a physiological autoregulatory approach. Nat Med 15: 447–454, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
40. Cao L, Liu X, Lin EJ, Wang C, Choi EY, Riban V, Lin B, During MJ. Environmental and genetic activation of a brain-adipocyte BDNF/leptin axis causes cancer remission and inhibition. Cell 142: 52–64, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
41. Carling D, Zammit VA, Hardie DG. A common bicyclic protein kinase cascade inactivates the regulatory enzymes of fatty acid and cholesterol biosynthesis. FEBS Lett 223: 217–222, 1987 [Abstract] [Google Scholar]
42. Cassiman D, Denef C, Desmet VJ, Roskams T. Human and rat hepatic stellate cells express neurotrophins and neurotrophin receptors. Hepatology 33: 148–158, 2001 [Abstract] [Google Scholar]
43. Castren E, Thoenen H, Lindholm D. Brain-derived neurotrophic factor messenger RNA is expressed in the septum, hypothalamus and in adrenergic brain stem nuclei of adult rat brain and is increased by osmotic stimulation in the paraventricular nucleus. Neuroscience 64: 71–80, 1995 [Abstract] [Google Scholar]
44. Cenquizca LA, Swanson LW. Analysis of direct hippocampal cortical field CA1 axonal projections to diencephalon in the rat. J Comp Neurol 497: 101–114, 2006 [Europe PMC free article] [Abstract] [Google Scholar]
45. Chakravarthy S, Saiepour MH, Bence M, Perry S, Hartman R, Couey JJ, Mansvelder HD, Levelt CN. Postsynaptic TrkB signaling has distinct roles in spine maintenance in adult visual cortex and hippocampus. Proc Natl Acad Sci USA 103: 1071–1076, 2006 [Europe PMC free article] [Abstract] [Google Scholar]
46. Chang GQ, Gaysinskaya V, Karatayev O, Leibowitz SF. Maternal high-fat diet and fetal programming: increased proliferation of hypothalamic peptide-producing neurons that increase risk for overeating and obesity. J Neurosci 28: 12107–12119, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
47. Chen MJ, Russo-Neustadt A. A running exercise-induced up-regulation of hippocampal brain-derived neurotrophic factor is CREB-dependent. Hippocampus 19: 962–972, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
48. Chen ZY, Patel PD, Sant G, Meng CX, Teng KK, Hempstead BL, Lee FS. Variant brain-derived neurotrophic factor (BDNF) (Met66) alters the intracellular trafficking and activity-dependent secretion of wild-type BDNF in neurosecretory cells and cortical neurons. J Neurosci 24: 4401–4411, 2004 [Abstract] [Google Scholar]
49. Clifton PG, Vickers SP, Somerville EM. Little and often: ingestive behavior patterns following hippocampal lesions in rats. Behav Neurosci 112: 502–511, 1998 [Abstract] [Google Scholar]
50. Coleman CG, Wang G, Park L, Anrather J, Delagrammatikas GJ, Chan J, Zhou J, Iadecola C, Pickel VM. Chronic intermittent hypoxia induces NMDA receptor-dependent plasticity and suppresses nitric oxide signaling in the mouse hypothalamic paraventricular nucleus. J Neurosci 30: 12103–12112, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
51. Conner JM, Lauterborn JC, Yan Q, Gall CM, Varon S. Distribution of brain-derived neurotrophic factor (BDNF) protein and mRNA in the normal adult rat CNS: evidence for anterograde axonal transport. J Neurosci 17: 2295–2313, 1997 [Abstract] [Google Scholar]
52. Cordeira JW, Frank L, Sena-Esteves M, Pothos EN, Rios M. Brain-derived neurotrophic factor regulates hedonic feeding by acting on the mesolimbic dopamine system. J Neurosci 30: 2533–2541, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
53. Cotman CW, Berchtold NC, Christie LA. Exercise builds brain health: key roles of growth factor cascades and inflammation. Trends Neurosci 30: 464–472, 2007 [Abstract] [Google Scholar]
54. Cunha C, Brambilla R, Thomas KL. A simple role for BDNF in learning and memory? Front Mol Neurosci 3: 1, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
55. Danzer SC, Kotloski RJ, Walter C, Hughes M, McNamara JO. Altered morphology of hippocampal dentate granule cell presynaptic and postsynaptic terminals following conditional deletion of TrkB. Hippocampus 18: 668–678, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
56. Dardennes RM, Zizzari P, Tolle V, Foulon C, Kipman A, Romo L, Iancu-Gontard D, Boni C, Sinet PM, Therese Bluet M, Estour B, Mouren MC, Guelfi JD, Rouillon F, Gorwood P, Epelbaum J. Family trios analysis of common polymorphisms in the obestatin/ghrelin, BDNF and AGRP genes in patients with Anorexia nervosa: association with subtype, body-mass index, severity and age of onset. Psychoneuroendocrinology 32: 106–113, 2007 [Abstract] [Google Scholar]
57. Davidson TL, Chan K, Jarrard LE, Kanoski SE, Clegg DJ, Benoit SC. Contributions of the hippocampus and medial prefrontal cortex to energy and body weight regulation. Hippocampus 19: 235–252, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
58. Davidson TL, Kanoski SE, Walls EK, Jarrard LE. Memory inhibition and energy regulation. Physiol Behav 86: 731–746, 2005 [Abstract] [Google Scholar]
59. Davidson TL, McKernan MG, Jarrard LE. Hippocampal lesions do not impair negative patterning: a challenge to configural association theory. Behav Neurosci 107: 227–234, 1993 [Abstract] [Google Scholar]
60. DelParigi A, Chen K, Salbe AD, Hill JO, Wing RR, Reiman EM, Tataranni PA. Persistence of abnormal neural responses to a meal in postobese individuals. Int J Obes Relat Metab Disord 28: 370–377, 2004 [Abstract] [Google Scholar]
61. Drake CT, Milner TA, Patterson SL. Ultrastructural localization of full-length trkB immunoreactivity in rat hippocampus suggests multiple roles in modulating activity-dependent synaptic plasticity. J Neurosci 19: 8009–8026, 1999 [Abstract] [Google Scholar]
62. Duan W, Guo Z, Jiang H, Ware M, Li XJ, Mattson MP. Dietary restriction normalizes glucose metabolism and BDNF levels, slows disease progression, and increases survival in huntingtin mutant mice. Proc Natl Acad Sci USA 100: 2911–2916, 2003 [Europe PMC free article] [Abstract] [Google Scholar]
63. During MJ, Cao L. VEGF, a mediator of the effect of experience on hippocampal neurogenesis. Curr Alzheimer Res 3: 29–33, 2006 [Abstract] [Google Scholar]
64. Egan MF, Kojima M, Callicott JH, Goldberg TE, Kolachana BS, Bertolino A, Zaitsev E, Gold B, Goldman D, Dean M, Lu B, Weinberger DR. The BDNF val66met polymorphism affects activity-dependent secretion of BDNF and human memory and hippocampal function. Cell 112: 257–269, 2003 [Abstract] [Google Scholar]
65. Eide FF, Vining ER, Eide BL, Zang K, Wang XY, Reichardt LF. Naturally occurring truncated trkB receptors have dominant inhibitory effects on brain-derived neurotrophic factor signaling. J Neurosci 16: 3123–3129, 1996 [Europe PMC free article] [Abstract] [Google Scholar]
66. El-Gharbawy AH, Adler-Wailes DC, Mirch MC, Theim KR, Ranzenhofer L, Tanofsky-Kraff M, Yanovski JA. Serum brain-derived neurotrophic factor concentrations in lean and overweight children and adolescents. J Clin Endocrinol Metab 91: 3548–3552, 2006 [Europe PMC free article] [Abstract] [Google Scholar]
67. Ernfors P, Ibanez CF, Ebendal T, Olson L, Persson H. Molecular cloning and neurotrophic activities of a protein with structural similarities to nerve growth factor: developmental and topographical expression in the brain. Proc Natl Acad Sci USA 87: 5454–5458, 1990 [Europe PMC free article] [Abstract] [Google Scholar]
68. Esterbauer H, Schaur RJ, Zollner H. Chemistry and biochemistry of 4-hydroxynonenal, malonaldehyde and related aldehydes. Free Radic Biol Med 11: 81–128, 1991 [Abstract] [Google Scholar]
69. Fahnestock M, Michalski B, Xu B, Coughlin MD. The precursor pro-nerve growth factor is the predominant form of nerve growth factor in brain and is increased in Alzheimer's disease. Mol Cell Neurosci 18: 210–220, 2001 [Abstract] [Google Scholar]
70. Fahrbach SE, Morrell JI, Pfaff DW. Studies of ventromedial hypothalamic afferents in the rat using three methods of HRP application. Exp Brain Res 77: 221–233, 1989 [Abstract] [Google Scholar]
71. Fekete C, Marks DL, Sarkar S, Emerson CH, Rand WM, Cone RD, Lechan RM. Effect of Agouti-related protein in regulation of the hypothalamic-pituitary-thyroid axis in the melanocortin 4 receptor knockout mouse. Endocrinology 145: 4816–4821, 2004 [Abstract] [Google Scholar]
72. Fekete C, Mihaly E, Luo LG, Kelly J, Clausen JT, Mao Q, Rand WM, Moss LG, Kuhar M, Emerson CH, Jackson IM, Lechan RM. Association of cocaine- and amphetamine-regulated transcript-immunoreactive elements with thyrotropin-releasing hormone-synthesizing neurons in the hypothalamic paraventricular nucleus and its role in the regulation of the hypothalamic-pituitary-thyroid axis during fasting. J Neurosci 20: 9224–9234, 2000 [Abstract] [Google Scholar]
73. Ferrer I, Blanco R, Cutillas B, Ambrosio S. Fas and Fas-L expression in Huntington's disease and Parkinson's disease. Neuropathol Appl Neurobiol 26: 424–433, 2000 [Abstract] [Google Scholar]
74. Ferris LT, Williams JS, Shen CL. The effect of acute exercise on serum brain-derived neurotrophic factor levels and cognitive function. Med Sci Sports Exerc 39: 728–734, 2007 [Abstract] [Google Scholar]
75. Figlewicz DP, MacDonald Naleid A, Sipols AJ. Modulation of food reward by adiposity signals. Physiol Behav 91: 473–478, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
76. Figurov A, Pozzo-Miller LD, Olafsson P, Wang T, Lu B. Regulation of synaptic responses to high-frequency stimulation and LTP by neurotrophins in the hippocampus. Nature 381: 706–709, 1996 [Abstract] [Google Scholar]
77. Fontan-Lozano A, Saez-Cassanelli JL, Inda MC, de los Santos-Arteaga M, Sierra-Dominguez SA, Lopez-Lluch G, Delgado-Garcia JM, Carrion AM. Caloric restriction increases learning consolidation and facilitates synaptic plasticity through mechanisms dependent on NR2B subunits of the NMDA receptor. J Neurosci 27: 10185–10195, 2007 [Abstract] [Google Scholar]
78. Friedman WJ. Neurotrophins induce death of hippocampal neurons via the p75 receptor. J Neurosci 20: 6340–6346, 2000 [Abstract] [Google Scholar]
79. Fruebis J, Tsao TS, Javorschi S, Ebbets-Reed D, Erickson MR, Yen FT, Bihain BE, Lodish HF. Proteolytic cleavage product of 30-kDa adipocyte complement-related protein increases fatty acid oxidation in muscle and causes weight loss in mice. Proc Natl Acad Sci USA 98: 2005–2010, 2001 [Europe PMC free article] [Abstract] [Google Scholar]
80. Fuchikami M, Morinobu S, Kurata A, Yamamoto S, Yamawaki S. Single immobilization stress differentially alters the expression profile of transcripts of the brain-derived neurotrophic factor (BDNF) gene and histone acetylation at its promoters in the rat hippocampus. Int J Neuropsychopharmacol 12: 73–82, 2009 [Abstract] [Google Scholar]
81. Fujimura H, Altar CA, Chen R, Nakamura T, Nakahashi T, Kambayashi J, Sun B, Tandon NN. Brain-derived neurotrophic factor is stored in human platelets and released by agonist stimulation. Thromb Haemost 87: 728–734, 2002 [Abstract] [Google Scholar]
82. Fujinami A, Ohta K, Obayashi H, Fukui M, Hasegawa G, Nakamura N, Kozai H, Imai S, Ohta M. Serum brain-derived neurotrophic factor in patients with type 2 diabetes mellitus: Relationship to glucose metabolism and biomarkers of insulin resistance. Clin Biochem 41: 812–817, 2008 [Abstract] [Google Scholar]
83. Gage FH, Kempermann G, Palmer TD, Peterson DA, Ray J. Multipotent progenitor cells in the adult dentate gyrus. J Neurobiol 36: 249–266, 1998 [Abstract] [Google Scholar]
84. Gelegen C, van den Heuvel J, Collier DA, Campbell IC, Oppelaar H, Hessel E, Kas MJ. Dopaminergic and brain-derived neurotrophic factor signalling in inbred mice exposed to a restricted feeding schedule. Genes Brain Behav 7: 552–559, 2008 [Abstract] [Google Scholar]
85. Gielen A, Khademi M, Muhallab S, Olsson T, Piehl F. Increased brain-derived neurotrophic factor expression in white blood cells of relapsing-remitting multiple sclerosis patients. Scand J Immunol 57: 493–497, 2003 [Abstract] [Google Scholar]
86. Givalois L, Arancibia S, Tapia-Arancibia L. Concomitant changes in CRH mRNA levels in rat hippocampus and hypothalamus following immobilization stress. Brain Res Mol Brain Res 75: 166–171, 2000 [Abstract] [Google Scholar]
87. Goodman LJ, Valverde J, Lim F, Geschwind MD, Federoff HJ, Geller AI, Hefti F. Regulated release and polarized localization of brain-derived neurotrophic factor in hippocampal neurons. Mol Cell Neurosci 7: 222–238, 1996 [Abstract] [Google Scholar]
88. Gotz R, Koster R, Winkler C, Raulf F, Lottspeich F, Schartl M, Thoenen H. Neurotrophin-6 is a new member of the nerve growth factor family. Nature 372: 266–269, 1994 [Abstract] [Google Scholar]
89. Gould E, Gross CG. Neurogenesis in adult mammals: some progress and problems. J Neurosci 22: 619–623, 2002 [Abstract] [Google Scholar]
90. Gratacos M, Gonzalez JR, Mercader JM, de Cid R, Urretavizcaya M, Estivill X. Brain-derived neurotrophic factor Val66Met and psychiatric disorders: meta-analysis of case-control studies confirm association to substance-related disorders, eating disorders, and schizophrenia. Biol Psychiatry 61: 911–922, 2007 [Abstract] [Google Scholar]
91. Gray J, Yeo GS, Cox JJ, Morton J, Adlam AL, Keogh JM, Yanovski JA, El Gharbawy A, Han JC, Tung YC, Hodges JR, Raymond FL, O'Rahilly S, Farooqi IS. 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: 3366–3371, 2006 [Europe PMC free article] [Abstract] [Google Scholar]
92. Greenwood CE, Winocur G. Cognitive impairment in rats fed high-fat diets: a specific effect of saturated fatty-acid intake. Behav Neurosci 110: 451–459, 1996 [Abstract] [Google Scholar]
93. Greenwood CE, Winocur G. High-fat diets, insulin resistance and declining cognitive function. Neurobiol Aging 26 Suppl 1: 42–45, 2005 [Abstract] [Google Scholar]
94. Greisen MH, Altar CA, Bolwig TG, Whitehead R, Wortwein G. Increased adult hippocampal brain-derived neurotrophic factor and normal levels of neurogenesis in maternal separation rats. J Neurosci Res 79: 772–778, 2005 [Abstract] [Google Scholar]
95. Griffin EW, Bechara RG, Birch AM, Kelly AM. Exercise enhances hippocampal-dependent learning in the rat: evidence for a BDNF-related mechanism. Hippocampus 19: 973–980, 2009 [Abstract] [Google Scholar]
96. Grill HJ, Kaplan JM. The neuroanatomical axis for control of energy balance. Front Neuroendocrinol 23: 2–40, 2002 [Abstract] [Google Scholar]
97. Grothe C, Unsicker K. Neuron-enriched cultures of adult rat dorsal root ganglia: establishment, characterization, survival, and neuropeptide expression in response to trophic factors. J Neurosci Res 18: 539–550, 1987 [Abstract] [Google Scholar]
98. Haapasalo A, Sipola I, Larsson K, Akerman KE, Stoilov P, Stamm S, Wong G, Castren E. Regulation of TRKB surface expression by brain-derived neurotrophic factor and truncated TRKB isoforms. J Biol Chem 277: 43160–43167, 2002 [Abstract] [Google Scholar]
99. Han JC, Liu QR, Jones M, Levinn RL, Menzie CM, Jefferson-George KS, Adler-Wailes DC, Sanford EL, Lacbawan FL, Uhl GR, Rennert OM, Yanovski JA. Brain-derived neurotrophic factor and obesity in the WAGR syndrome. N Engl J Med 359: 918–927, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
100. Hartmann M, Heumann R, Lessmann V. Synaptic secretion of BDNF after high-frequency stimulation of glutamatergic synapses. EMBO J 20: 5887–5897, 2001 [Europe PMC free article] [Abstract] [Google Scholar]
101. Haskell-Luevano C, Chen P, Li C, Chang K, Smith MS, Cameron JL, Cone RD. Characterization of the neuroanatomical distribution of agouti-related protein immunoreactivity in the rhesus monkey and the rat. Endocrinology 140: 1408–1415, 1999 [Abstract] [Google Scholar]
102. Hebben N, Corkin S, Eichenbaum H, Shedlack K. Diminished ability to interpret and report internal states after bilateral medial temporal resection: case HM. Behav Neurosci 99: 1031–1039, 1985 [Abstract] [Google Scholar]
103. Hewitt SA, Bains JS. Brain-derived neurotrophic factor silences GABA synapses onto hypothalamic neuroendocrine cells through a postsynaptic dynamin-mediated mechanism. J Neurophysiol 95: 2193–2198, 2006 [Abstract] [Google Scholar]
104. Hock CH, Heese K, Olivieri G, Hulette CH, Rosenberg C, Nitsch RM, Otten U. Alterations in neurotrophins and neurotrophin receptors in Alzheimer's disease. J Neural Transm Suppl 59: 171–174, 2000 [Abstract] [Google Scholar]
105. Hofer MM, Barde YA. Brain-derived neurotrophic factor prevents neuronal death in vivo. Nature 331: 261–262, 1988 [Abstract] [Google Scholar]
106. Houmard JA. Intramuscular lipid oxidation and obesity. Am J Physiol Regul Integr Comp Physiol 294: R1111–R1116, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
107. Hulver MW, Dohm GL. The molecular mechanism linking muscle fat accumulation to insulin resistance. Proc Nutr Soc 63: 375–380, 2004 [Abstract] [Google Scholar]
108. Huszar D, Lynch CA, Fairchild-Huntress V, Dunmore JH, Fang Q, Berkemeier LR, Gu W, Kesterson RA, Boston BA, Cone RD, Smith FJ, Campfield LA, Burn P, Lee F. Targeted disruption of the melanocortin-4 receptor results in obesity in mice. Cell 88: 131–141, 1997 [Abstract] [Google Scholar]
109. Ikeda Y, Luo X, Abbud R, Nilson JH, Parker KL. The nuclear receptor steroidogenic factor 1 is essential for the formation of the ventromedial hypothalamic nucleus. Mol Endocrinol 9: 478–486, 1995 [Abstract] [Google Scholar]
110. Jarrard LE. What does the hippocampus really do? Behav Brain Res 71: 1–10, 1995 [Abstract] [Google Scholar]
111. Jean A. Brain stem control of swallowing: neuronal network and cellular mechanisms. Physiol Rev 81: 929–969, 2001 [Abstract] [Google Scholar]
112. Jones DP. Redefining oxidative stress. Antioxid Redox Signal 8: 1865–1879, 2006 [Abstract] [Google Scholar]
113. Joosten EA, Houweling DA. Local acute application of BDNF in the lesioned spinal cord anti-inflammatory and anti-oxidant effects. Neuroreport 15: 1163–1166, 2004 [Abstract] [Google Scholar]
114. Jovanovic JN, Benfenati F, Siow YL, Sihra TS, Sanghera JS, Pelech SL, Greengard P, Czernik AJ. Neurotrophins stimulate phosphorylation of synapsin I by MAP kinase and regulate synapsin I-actin interactions. Proc Natl Acad Sci USA 93: 3679–3683, 1996 [Europe PMC free article] [Abstract] [Google Scholar]
115. Ka S, Lindberg J, Stromstedt L, Fitzsimmons C, Lindqvist N, Lundeberg J, Siegel PB, Andersson L, Hallbook F. Extremely different behaviours in high and low body weight lines of chicken are associated with differential expression of genes involved in neuronal plasticity. J Neuroendocrinol 21: 208–216, 2009 [Abstract] [Google Scholar]
116. Kalcheim C, Gendreau M. Brain-derived neurotrophic factor stimulates survival and neuronal differentiation in cultured avian neural crest. Brain Res 469: 79–86, 1988 [Abstract] [Google Scholar]
117. Kang H, Welcher AA, Shelton D, Schuman EM. Neurotrophins and time: different roles for TrkB signaling in hippocampal long-term potentiation. Neuron 19: 653–664, 1997 [Abstract] [Google Scholar]
118. Kanoski SE, Meisel RL, Mullins AJ, Davidson TL. The effects of energy-rich diets on discrimination reversal learning and on BDNF in the hippocampus and prefrontal cortex of the rat. Behav Brain Res 182: 57–66, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
119. Katoh-Semba R, Takeuchi IK, Semba R, Kato K. Distribution of brain-derived neurotrophic factor in rats and its changes with development in the brain. J Neurochem 69: 34–42, 1997 [Abstract] [Google Scholar]
120. Katz A, Meiri N. Brain-derived neurotrophic factor is critically involved in thermal-experience-dependent developmental plasticity. J Neurosci 26: 3899–3907, 2006 [Abstract] [Google Scholar]
121. Keller JN, Schmitt FA, Scheff SW, Ding Q, Chen Q, Butterfield DA, Markesbery WR. Evidence of increased oxidative damage in subjects with mild cognitive impairment. Neurology 64: 1152–1156, 2005 [Abstract] [Google Scholar]
122. Kelley AE. Ventral striatal control of appetitive motivation: role in ingestive behavior and reward-related learning. Neurosci Biobehav Rev 27: 765–776, 2004 [Abstract] [Google Scholar]
123. Kernie SG, Liebl DJ, Parada LF. BDNF regulates eating behavior and locomotor activity in mice. EMBO J 19: 1290–1300, 2000 [Europe PMC free article] [Abstract] [Google Scholar]
124. Kerschensteiner M, Gallmeier E, Behrens L, Leal VV, Misgeld T, Klinkert WE, Kolbeck R, Hoppe E, Oropeza-Wekerle RL, Bartke I, Stadelmann C, Lassmann H, Wekerle H, Hohlfeld R. Activated human T cells, B cells, and monocytes produce brain-derived neurotrophic factor in vitro and in inflammatory brain lesions: a neuroprotective role of inflammation? J Exp Med 189: 865–870, 1999 [Europe PMC free article] [Abstract] [Google Scholar]
125. Kharitonenkov A, Shiyanova TL, Koester A, Ford AM, Micanovic R, Galbreath EJ, Sandusky GE, Hammond LJ, Moyers JS, Owens RA, Gromada J, Brozinick JT, Hawkins ED, Wroblewski VJ, Li DS, Mehrbod F, Jaskunas SR, Shanafelt AB. FGF-21 as a novel metabolic regulator. J Clin Invest 115: 1627–1635, 2005 [Abstract] [Google Scholar]
126. King BM. The rise, fall, and resurrection of the ventromedial hypothalamus in the regulation of feeding behavior and body weight. Physiol Behav 87: 221–244, 2006 [Abstract] [Google Scholar]
127. King BM, Phelps GR, Frohman LA. Hypothalamic obesity in female rats in absence of vagally mediated hyperinsulinemia. Am J Physiol Endocrinol Metab 239: E437–E441, 1980 [Abstract] [Google Scholar]
128. King BM, Rossiter KN, Stines SG, Zaharan GM, Cook JT, Humphries MD, York DA. Amygdaloid-lesion hyperphagia: impaired response to caloric challenges and altered macronutrient selection. Am J Physiol Regul Integr Comp Physiol 275: R485–R493, 1998 [Abstract] [Google Scholar]
129. Kitazawa H, Numakawa T, Adachi N, Kumamaru E, Tuerxun T, Kudo M, Kunugi H. Cyclophosphamide promotes cell survival via activation of intracellular signaling in cultured cortical neurons. Neurosci Lett 470: 139–144, 2010 [Abstract] [Google Scholar]
130. Klein R, Nanduri V, Jing SA, Lamballe F, Tapley P, Bryant S, Cordon-Cardo C, Jones KR, Reichardt LF, Barbacid M. The trkB tyrosine protein kinase is a receptor for brain-derived neurotrophic factor and neurotrophin-3. Cell 66: 395–403, 1991 [Europe PMC free article] [Abstract] [Google Scholar]
131. Knusel B, Hefti F. K-252b is a selective and nontoxic inhibitor of nerve growth factor action on cultured brain neurons. J Neurochem 57: 955–962, 1991 [Abstract] [Google Scholar]
132. Kohjima M, Sun Y, Chan L. Increased food intake leads to obesity and insulin resistance in the tg2576 Alzheimer's disease mouse model. Endocrinology 151: 1532–1540, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
133. Kokaia Z, Nawa H, Uchino H, Elmer E, Kokaia M, Carnahan J, Smith ML, Siesjo BK, Lindvall O. Regional brain-derived neurotrophic factor mRNA and protein levels following transient forebrain ischemia in the rat. Brain Res Mol Brain Res 38: 139–144, 1996 [Abstract] [Google Scholar]
134. Kokoeva MV, Yin H, Flier JS. Evidence for constitutive neural cell proliferation in the adult murine hypothalamus. J Comp Neurol 505: 209–220, 2007 [Abstract] [Google Scholar]
135. Kokoeva MV, Yin H, Flier JS. Neurogenesis in the hypothalamus of adult mice: potential role in energy balance. Science 310: 679–683, 2005 [Abstract] [Google Scholar]
136. Kolbeck R, Jungbluth S, Barde YA. Characterisation of neurotrophin dimers and monomers. Eur J Biochem 225: 995–1003, 1994 [Abstract] [Google Scholar]
137. Komori T, Morikawa Y, Nanjo K, Senba E. Induction of brain-derived neurotrophic factor by leptin in the ventromedial hypothalamus. Neuroscience 139: 1107–1115, 2006 [Abstract] [Google Scholar]
138. Kong WM, Martin NM, Smith KL, Gardiner JV, Connoley IP, Stephens DA, Dhillo WS, Ghatei MA, Small CJ, Bloom SR. Triiodothyronine stimulates food intake via the hypothalamic ventromedial nucleus independent of changes in energy expenditure. Endocrinology 145: 5252–5258, 2004 [Abstract] [Google Scholar]
139. Korte M, Carroll P, Wolf E, Brem G, Thoenen H, Bonhoeffer T. Hippocampal long-term potentiation is impaired in mice lacking brain-derived neurotrophic factor. Proc Natl Acad Sci USA 92: 8856–8860, 1995 [Europe PMC free article] [Abstract] [Google Scholar]
140. Korte M, Staiger V, Griesbeck O, Thoenen H, Bonhoeffer T. The involvement of brain-derived neurotrophic factor in hippocampal long-term potentiation revealed by gene targeting experiments. J Physiol (Paris) 90: 157–164, 1996 [Abstract] [Google Scholar]
141. Krabbe KS, Nielsen AR, Krogh-Madsen R, Plomgaard P, Rasmussen P, Erikstrup C, Fischer CP, Lindegaard B, Petersen AM, Taudorf S, Secher NH, Pilegaard H, Bruunsgaard H, Pedersen BK. Brain-derived neurotrophic factor (BDNF) and type 2 diabetes. Diabetologia 50: 431–438, 2007 [Abstract] [Google Scholar]
142. Kumar S, Parkash J, Kataria H, Kaur G. Interactive effect of excitotoxic injury and dietary restriction on neurogenesis and neurotrophic factors in adult male rat brain. Neurosci Res 65: 367–374, 2009 [Abstract] [Google Scholar]
143. Lafenetre P, Leske O, Ma-Hogemeie Z, Haghikia A, Bichler Z, Wahle P, Heumann R. Exercise can rescue recognition memory impairment in a model with reduced adult hippocampal neurogenesis (Online). Front Behav Neurosci 3: 34, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
144. Lambert PD, Anderson KD, Sleeman MW, Wong V, Tan J, Hijarunguru A, Corcoran TL, Murray JD, Thabet KE, Yancopoulos GD, Wiegand SJ. Ciliary neurotrophic factor activates leptin-like pathways and reduces body fat, without cachexia or rebound weight gain, even in leptin-resistant obesity. Proc Natl Acad Sci USA 98: 4652–4657, 2001 [Europe PMC free article] [Abstract] [Google Scholar]
145. Lapchak PA, Hefti F. BDNF and NGF treatment in lesioned rats: effects on cholinergic function and weight gain. Neuroreport 3: 405–408, 1992 [Abstract] [Google Scholar]
146. Larkfors L, Lindsay RM, Alderson RF. Ciliary neurotrophic factor enhances the survival of Purkinje cells in vitro. Eur J Neurosci 6: 1015–1025, 1994 [Abstract] [Google Scholar]
147. Larsson E, Nanobashvili A, Kokaia Z, Lindvall O. Evidence for neuroprotective effects of endogenous brain-derived neurotrophic factor after global forebrain ischemia in rats. J Cereb Blood Flow Metab 19: 1220–1228, 1999 [Abstract] [Google Scholar]
148. Lebrun B, Bariohay B, Moyse E, Jean A. Brain-derived neurotrophic factor (BDNF) and food intake regulation: a minireview. Auton Neurosci 126–127: 30–38, 2006 [Abstract] [Google Scholar]
149. Lechan RM, Fekete C. Central mechanisms for thyroid hormone regulation (Online). Am J Psychiatry 163: 1492, 2006 [Abstract] [Google Scholar]
150. Lee J, Seroogy KB, Mattson MP. Dietary restriction enhances neurotrophin expression and neurogenesis in the hippocampus of adult mice. J Neurochem 80: 539–547, 2002 [Abstract] [Google Scholar]
151. Leibrock J, Lottspeich F, Hohn A, Hofer M, Hengerer B, Masiakowski P, Thoenen H, Barde YA. Molecular cloning and expression of brain-derived neurotrophic factor. Nature 341: 149–152, 1989 [Abstract] [Google Scholar]
152. Levi-Montalcini R. The nerve growth factor: thirty-five years later. Biosci Rep 7: 681–699, 1987 [Abstract] [Google Scholar]
153. Li DP, Yang Q, Pan HM, Pan HL. Plasticity of pre- and postsynaptic GABAB receptor function in the paraventricular nucleus in spontaneously hypertensive rats. Am J Physiol Heart Circ Physiol 295: H807–H815, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
154. Li W, Keifer J. BDNF-induced synaptic delivery of AMPAR subunits is differentially dependent on NMDA receptors and requires ERK. Neurobiol Learn Mem 91: 243–249, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
155. Li W, Keifer J. Coordinate action of pre- and postsynaptic brain-derived neurotrophic factor is required for AMPAR trafficking and acquisition of in vitro classical conditioning. Neuroscience 155: 686–697, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
156. Lin JC, Tsao D, Barras P, Bastarrachea RA, Boyd B, Chou J, Rosete R, Long H, Forgie A, Abdiche Y, Dilley J, Stratton J, Garcia C, Sloane DL, Comuzzie AG, Rosenthal A. Appetite enhancement and weight gain by peripheral administration of TrkB agonists in non-human primates. PLoS One 3: e1900, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
157. Lin L, York DA. Amygdala enterostatin induces c-Fos expression in regions of hypothalamus that innervate the PVN. Brain Res 1020: 147–153, 2004 [Abstract] [Google Scholar]
158. Linthorst AC, Flachskamm C, Hopkins SJ, Hoadley ME, Labeur MS, Holsboer F, Reul JM. Long-term intracerebroventricular infusion of corticotropin-releasing hormone alters neuroendocrine, neurochemical, autonomic, behavioral, and cytokine responses to a systemic inflammatory challenge. J Neurosci 17: 4448–4460, 1997 [Abstract] [Google Scholar]
159. Lommatzsch M, Braun A, Mannsfeldt A, Botchkarev VA, Botchkareva NV, Paus R, Fischer A, Lewin GR, Renz H. Abundant production of brain-derived neurotrophic factor by adult visceral epithelia. Implications for paracrine and target-derived neurotrophic functions. Am J Pathol 155: 1183–1193, 1999 [Europe PMC free article] [Abstract] [Google Scholar]
160. Lu B. Pro-region of neurotrophins: role in synaptic modulation. Neuron 39: 735–738, 2003 [Abstract] [Google Scholar]
161. Luiten PG, Ono T, Nishijo H, Fukuda M. Differential input from the amygdaloid body to the ventromedial hypothalamic nucleus in the rat. Neurosci Lett 35: 253–258, 1983 [Abstract] [Google Scholar]
162. Luiten PG, Room P. Interrelations between lateral, dorsomedial and ventromedial hypothalamic nuclei in the rat. An HRP study. Brain Res 190: 321–332, 1980 [Abstract] [Google Scholar]
163. Maisonpierre PC, Belluscio L, Squinto S, Ip NY, Furth ME, Lindsay RM, Yancopoulos GD. Neurotrophin-3: a neurotrophic factor related to NGF and BDNF. Science 247: 1446–1451, 1990 [Abstract] [Google Scholar]
164. Malter JS. Regulation of mRNA stability in the nervous system and beyond. J Neurosci Res 66: 311–316, 2001 [Abstract] [Google Scholar]
165. Mandyam CD, Harburg GC, Eisch AJ. Determination of key aspects of precursor cell proliferation, cell cycle length and kinetics in the adult mouse subgranular zone. Neuroscience 146: 108–122, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
166. Maniam J, Morris MJ. Voluntary exercise and palatable high-fat diet both improve behavioural profile and stress responses in male rats exposed to early life stress: role of hippocampus. Psychoneuroendocrinology 35: 1553–1564, 2010 [Abstract] [Google Scholar]
167. Marmigere F, Givalois L, Rage F, Arancibia S, Tapia-Arancibia L. Rapid induction of BDNF expression in the hippocampus during immobilization stress challenge in adult rats. Hippocampus 13: 646–655, 2003 [Abstract] [Google Scholar]
168. Marmigere F, Rage F, Tapia-Arancibia L, Arancibia S. Expression of mRNAs encoding BDNF and its receptor in adult rat hypothalamus. Neuroreport 9: 1159–1163, 1998 [Abstract] [Google Scholar]
169. Martinez-Marcos A, Lanuza E, Halpern M. Organization of the ophidian amygdala: chemosensory pathways to the hypothalamus. J Comp Neurol 412: 51–68, 1999 [Abstract] [Google Scholar]
170. Matthews VB, Astrom MB, Chan MH, Bruce CR, Krabbe KS, Prelovsek O, Akerstrom T, Yfanti C, Broholm C, Mortensen OH, Penkowa M, Hojman P, Zankari A, Watt MJ, Bruunsgaard H, Pedersen BK, Febbraio MA. Brain-derived neurotrophic factor is produced by skeletal muscle cells in response to contraction and enhances fat oxidation via activation of AMP-activated protein kinase. Diabetologia 52: 1409–1418, 2009 [Abstract] [Google Scholar]
171. Mattson MP, Lovell MA, Furukawa K, Markesbery WR. Neurotrophic factors attenuate glutamate-induced accumulation of peroxides, elevation of intracellular Ca2+ concentration, and neurotoxicity and increase antioxidant enzyme activities in hippocampal neurons. J Neurochem 65: 1740–1751, 1995 [Abstract] [Google Scholar]
172. McGarry JD, Brown NF. The mitochondrial carnitine palmitoyltransferase system. From concept to molecular analysis. Eur J Biochem 244: 1–14, 1997 [Abstract] [Google Scholar]
173. McGarry JD, Takabayashi Y, Foster DW. The role of malonyl-coa in the coordination of fatty acid synthesis and oxidation in isolated rat hepatocytes. J Biol Chem 253: 8294–8300, 1978 [Abstract] [Google Scholar]
174. Mercader JM, Fernandez-Aranda F, Gratacos M, Ribases M, Badia A, Villarejo C, Solano R, Gonzalez JR, Vallejo J, Estivill X. Blood levels of brain-derived neurotrophic factor correlate with several psychopathological symptoms in anorexia nervosa patients. Neuropsychobiology 56: 185–190, 2007 [Abstract] [Google Scholar]
175. Merhi ZO, Minkoff H, Lambert-Messerlian GM, Macura J, Feldman J, Seifer DB. Plasma brain-derived neurotrophic factor in women after bariatric surgery: a pilot study. Fertil Steril 91: 1544–1548, 2009 [Abstract] [Google Scholar]
176. Merlio JP, Ernfors P, Jaber M, Persson H. Molecular cloning of rat trkC and distribution of cells expressing messenger RNAs for members of the trk family in the rat central nervous system. Neuroscience 51: 513–532, 1992 [Abstract] [Google Scholar]
177. Molteni R, Barnard RJ, Ying Z, Roberts CK, Gomez-Pinilla F. A high-fat, refined sugar diet reduces hippocampal brain-derived neurotrophic factor, neuronal plasticity, and learning. Neuroscience 112: 803–814, 2002 [Abstract] [Google Scholar]
178. Molteni R, Wu A, Vaynman S, Ying Z, Barnard RJ, Gomez-Pinilla F. Exercise reverses the harmful effects of consumption of a high-fat diet on synaptic and behavioral plasticity associated to the action of brain-derived neurotrophic factor. Neuroscience 123: 429–440, 2004 [Abstract] [Google Scholar]
179. Moraes JC, Coope A, Morari J, Cintra DE, Roman EA, Pauli JR, Romanatto T, Carvalheira JB, Oliveira AL, Saad MJ, Velloso LA. High-fat diet induces apoptosis of hypothalamic neurons [Online]. PLoS One 4: e5045, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
180. Moser MB, Moser EI. Functional differentiation in the hippocampus. Hippocampus 8: 608–619, 1998 [Abstract] [Google Scholar]
181. Mountjoy KG, Mortrud MT, Low MJ, Simerly RB, Cone RD. Localization of the melanocortin-4 receptor (MC4-R) in neuroendocrine and autonomic control circuits in the brain. Mol Endocrinol 8: 1298–1308, 1994 [Abstract] [Google Scholar]
182. Mousavi K, Jasmin BJ. BDNF is expressed in skeletal muscle satellite cells and inhibits myogenic differentiation. J Neurosci 26: 5739–5749, 2006 [Abstract] [Google Scholar]
183. Mowla SJ, Farhadi HF, Pareek S, Atwal JK, Morris SJ, Seidah NG, Murphy RA. Biosynthesis and post-translational processing of the precursor to brain-derived neurotrophic factor. J Biol Chem 276: 12660–12666, 2001 [Abstract] [Google Scholar]
184. Mufson EJ, Kroin JS, Liu YT, Sobreviela T, Penn RD, Miller JA, Kordower JH. Intrastriatal and intraventricular infusion of brain-derived neurotrophic factor in the cynomologous monkey: distribution, retrograde transport and co-localization with substantia nigra dopamine-containing neurons. Neuroscience 71: 179–191, 1996 [Abstract] [Google Scholar]
185. Mufson EJ, Kroin JS, Sobreviela T, Burke MA, Kordower JH, Penn RD, Miller JA. Intrastriatal infusions of brain-derived neurotrophic factor: retrograde transport and colocalization with dopamine containing substantia nigra neurons in rat. Exp Neurol 129: 15–26, 1994 [Abstract] [Google Scholar]
186. Naert G, Ixart G, Tapia-Arancibia L, Givalois L. Continuous icv infusion of brain-derived neurotrophic factor modifies hypothalamic-pituitary-adrenal axis activity, locomotor activity and body temperature rhythms in adult male rats. Neuroscience 139: 779–789, 2006 [Abstract] [Google Scholar]
187. Nakagawa T, Ogawa Y, Ebihara K, Yamanaka M, Tsuchida A, Taiji M, Noguchi H, Nakao K. Anti-obesity and anti-diabetic effects of brain-derived neurotrophic factor in rodent models of leptin resistance. Int J Obes Relat Metab Disord 27: 557–565, 2003 [Abstract] [Google Scholar]
188. Nakagawa T, Tsuchida A, Itakura Y, Nonomura T, Ono M, Hirota F, Inoue T, Nakayama C, Taiji M, Noguchi H. Brain-derived neurotrophic factor regulates glucose metabolism by modulating energy balance in diabetic mice. Diabetes 49: 436–444, 2000 [Abstract] [Google Scholar]
189. Nakahashi T, Fujimura H, Altar CA, Li J, Kambayashi J, Tandon NN, Sun B. Vascular endothelial cells synthesize and secrete brain-derived neurotrophic factor. FEBS Lett 470: 113–117, 2000 [Abstract] [Google Scholar]
190. Nakatomi H, Kuriu T, Okabe S, Yamamoto S, Hatano O, Kawahara N, Tamura A, Kirino T, Nakafuku M. Regeneration of hippocampal pyramidal neurons after ischemic brain injury by recruitment of endogenous neural progenitors. Cell 110: 429–441, 2002 [Abstract] [Google Scholar]
191. Nakazato M, Tchanturia K, Schmidt U, Campbell IC, Treasure J, Collier DA, Hashimoto K, Iyo M. Brain-derived neurotrophic factor (BDNF) and set-shifting in currently ill and recovered anorexia nervosa (AN) patients. Psychol Med 39: 1029–1035, 2009 [Abstract] [Google Scholar]
192. Narisawa-Saito M, Wakabayashi K, Tsuji S, Takahashi H, Nawa H. Regional specificity of alterations in NGF, BDNF and NT-3 levels in Alzheimer's disease. Neuroreport 7: 2925–2928, 1996 [Abstract] [Google Scholar]
193. Nawa H, Carnahan J, Gall C. BDNF protein measured by a novel enzyme immunoassay in normal brain and after seizure: partial disagreement with mRNA levels. Eur J Neurosci 7: 1527–1535, 1995 [Abstract] [Google Scholar]
194. Neeper SA, Gomez-Pinilla F, Choi J, Cotman CW. Physical activity increases mRNA for brain-derived neurotrophic factor and nerve growth factor in rat brain. Brain Res 726: 49–56, 1996 [Abstract] [Google Scholar]
195. Nichol K, Deeny SP, Seif J, Camaclang K, Cotman CW. Exercise improves cognition and hippocampal plasticity in APOE epsilon4 mice. Alzheimers Dement 5: 287–294, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
196. Nicholson JR, Peter JC, Lecourt AC, Barde YA, Hofbauer KG. Melanocortin-4 receptor activation stimulates hypothalamic brain-derived neurotrophic factor release to regulate food intake, body temperature and cardiovascular function. J Neuroendocrinol 19: 974–982, 2007 [Abstract] [Google Scholar]
197. Nonomura T, Tsuchida A, Ono-Kishino M, Nakagawa T, Taiji M, Noguchi H. Brain-derived neurotrophic factor regulates energy expenditure through the central nervous system in obese diabetic mice. Int J Exp Diabetes Res 2: 201–209, 2001 [Europe PMC free article] [Abstract] [Google Scholar]
198. Numan S, Seroogy KB. Expression of trkB and trkC mRNAs by adult midbrain dopamine neurons: a double-label in situ hybridization study. J Comp Neurol 403: 295–308, 1999 [Abstract] [Google Scholar]
199. O'Rahilly S, Yeo GS, Farooqi IS. Melanocortin receptors weigh in. Nat Med 10: 351–352, 2004 [Abstract] [Google Scholar]
200. Olson AK, Eadie BD, Ernst C, Christie BR. Environmental enrichment and voluntary exercise massively increase neurogenesis in the adult hippocampus via dissociable pathways. Hippocampus 16: 250–260, 2006 [Abstract] [Google Scholar]
201. Ong C, Han JC, Hyde TM, Lipska BK, Mou Z, Tsao JW, Kleinman JE, Yanovski JA. Associations of single nucleotide polymorphisms (SNPs) in brain-derived neurotrophic factor (BDNF) with BDNF expression in human ventromedial hypothalamus (VMH) and with body mass index. Obesity 18: S61–S62, 2010 [Google Scholar]
202. Pang PT, Teng HK, Zaitsev E, Woo NT, Sakata K, Zhen S, Teng KK, Yung WH, Hempstead BL, Lu B. Cleavage of proBDNF by tPA/plasmin is essential for long-term hippocampal plasticity. Science 306: 487–491, 2004 [Abstract] [Google Scholar]
203. Pardridge WM, Kang YS, Buciak JL. Transport of human recombinant brain-derived neurotrophic factor (BDNF) through the rat blood-brain barrier in vivo using vector-mediated peptide drug delivery. Pharm Res 11: 738–746, 1994 [Abstract] [Google Scholar]
204. Park HR, Park M, Choi J, Park KY, Chung HY, Lee J. A high-fat diet impairs neurogenesis: involvement of lipid peroxidation and brain-derived neurotrophic factor. Neurosci Lett 482: 235–239, 2010 [Abstract] [Google Scholar]
205. Patterson SL, Abel T, Deuel TA, Martin KC, Rose JC, Kandel ER. Recombinant BDNF rescues deficits in basal synaptic transmission and hippocampal LTP in BDNF knockout mice. Neuron 16: 1137–1145, 1996 [Abstract] [Google Scholar]
206. Paxinos G, Watson C. The Rat Brain in Stereotaxic Coordinates. Amsterdam, The Netherlands: Elsevier Academic, 2007 [Google Scholar]
207. Pelleymounter MA, Cullen MJ, Wellman CL. Characteristics of BDNF-induced weight loss. Exp Neurol 131: 229–238, 1995 [Abstract] [Google Scholar]
208. Pencea V, Bingaman KD, Wiegand SJ, Luskin MB. Infusion of brain-derived neurotrophic factor into the lateral ventricle of the adult rat leads to new neurons in the parenchyma of the striatum, septum, thalamus, and hypothalamus. J Neurosci 21: 6706–6717, 2001 [Abstract] [Google Scholar]
209. Peng CH, Chiou SH, Chen SJ, Chou YC, Ku HH, Cheng CK, Yen CJ, Tsai TH, Chang YL, Kao CL. Neuroprotection by imipramine against lipopolysaccharide-induced apoptosis in hippocampus-derived neural stem cells mediated by activation of BDNF and the MAPK pathway. Eur Neuropsychopharmacol 18: 128–140, 2008 [Abstract] [Google Scholar]
210. Porter DW, Kerr BD, Flatt PR, Holscher C, Gault VA. Four weeks administration of Liraglutide improves memory and learning as well as glycaemic control in mice with high fat dietary-induced obesity and insulin resistance. Diabetes Obes Metab 12: 891–899, 2010 [Abstract] [Google Scholar]
211. Pruunsild P, Kazantseva A, Aid T, Palm K, Timmusk T. Dissecting the human BDNF locus: bidirectional transcription, complex splicing, and multiple promoters. Genomics 90: 397–406, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
212. Radziejewski C, Robinson RC, DiStefano PS, Taylor JW. Dimeric structure and conformational stability of brain-derived neurotrophic factor and neurotrophin-3. Biochemistry 31: 4431–4436, 1992 [Abstract] [Google Scholar]
213. Rage F, Givalois L, Marmigere F, Tapia-Arancibia L, Arancibia S. Immobilization stress rapidly modulates BDNF mRNA expression in the hypothalamus of adult male rats. Neuroscience 112: 309–318, 2002 [Abstract] [Google Scholar]
214. Ribases M, Gratacos M, Armengol L, de Cid R, Badia A, Jimenez L, Solano R, Vallejo J, Fernandez F, Estivill X. Met66 in the brain-derived neurotrophic factor (BDNF) precursor is associated with anorexia nervosa restrictive type. Mol Psychiatry 8: 745–751, 2003 [Abstract] [Google Scholar]
215. Ringman JM, Younkin SG, Pratico D, Seltzer W, Cole GM, Geschwind DH, Rodriguez-Agudelo Y, Schaffer B, Fein J, Sokolow S, Rosario ER, Gylys KH, Varpetian A, Medina LD, Cummings JL. Biochemical markers in persons with preclinical familial Alzheimer disease. Neurology 71: 85–92, 2008 [Abstract] [Google Scholar]
216. Rios M, Fan G, Fekete C, Kelly J, Bates B, Kuehn R, Lechan RM, Jaenisch R. Conditional deletion of brain-derived neurotrophic factor in the postnatal brain leads to obesity and hyperactivity. Mol Endocrinol 15: 1748–1757, 2001 [Abstract] [Google Scholar]
217. Rowsey PJ, Kluger MJ. Corticotropin releasing hormone is involved in exercise-induced elevation in core temperature. Psychoneuroendocrinology 19: 179–187, 1994 [Abstract] [Google Scholar]
218. Rozin P, Dow S, Moscovitch M, Rajaram S. What causes humans to begin and end a meal? A role for memory for what has been eaten, as evidenced by a study of multiple meal eating in amnesic patients. Psychol Sci 9: 392–396, 1998 [Google Scholar]
219. Saito S, Watanabe K, Hashimoto E, Saito T. Low serum BDNF and food intake regulation: a possible new explanation of the pathophysiology of eating disorders. Prog Neuropsychopharmacol Biol Psychiatry 33: 312–316, 2009 [Abstract] [Google Scholar]
220. Sakaguchi T, Arase K, Bray GA. Sympathetic activity and food intake of rats with ventromedial hypothalamic lesions. Int J Obes 12: 285–291, 1988 [Abstract] [Google Scholar]
221. Saper CB, Swanson LW, Cowan WM. The efferent connections of the ventromedial nucleus of the hypothalamus of the rat. J Comp Neurol 169: 409–442, 1976 [Abstract] [Google Scholar]
222. Savage DB, Petersen KF, Shulman GI. Disordered lipid metabolism and the pathogenesis of insulin resistance. Physiol Rev 87: 507–520, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
223. Schmelzeis MC, Mittleman G. The hippocampus and reward: effects of hippocampal lesions on progressive-ratio responding. Behav Neurosci 110: 1049–1066, 1996 [Abstract] [Google Scholar]
224. Sclafani A, Berner CN, Maul G. Multiple knife cuts between the medial and lateral hypothalamus in the rat: a reevaluation of hypothalamic feeding circuitry. J Comp Physiol Psychol 88: 201–207, 1975 [Abstract] [Google Scholar]
225. Seeley RJ, Burklow ML, Wilmer KA, Matthews CC, Reizes O, McOsker CC, Trokhan DP, Gross MC, Sheldon RJ. The effect of the melanocortin agonist, MT-II, on the defended level of body adiposity. Endocrinology 146: 3732–3738, 2005 [Abstract] [Google Scholar]
226. Seidah NG, Benjannet S, Pareek S, Chretien M, Murphy RA. Cellular processing of the neurotrophin precursors of NT3 and BDNF by the mammalian proprotein convertases. FEBS Lett 379: 247–250, 1996 [Abstract] [Google Scholar]
227. Seifert T, Brassard P, Wissenberg M, Rasmussen P, Nordby P, Stallknecht B, Adser H, Jakobsen AH, Pilegaard H, Nielsen HB, Secher NH. Endurance training enhances BDNF release from the human brain. Am J Physiol Regul Integr Comp Physiol 298: R372–R377, 2010 [Abstract] [Google Scholar]
228. Sendtner M, Schmalbruch H, Stockli KA, Carroll P, Kreutzberg GW, Thoenen H. Ciliary neurotrophic factor prevents degeneration of motor neurons in mouse mutant progressive motor neuronopathy. Nature 358: 502–504, 1992 [Abstract] [Google Scholar]
229. Seroogy KB, Lundgren KH, Tran TM, Guthrie KM, Isackson PJ, Gall CM. Dopaminergic neurons in rat ventral midbrain express brain-derived neurotrophic factor and neurotrophin-3 mRNAs. J Comp Neurol 342: 321–334, 1994 [Abstract] [Google Scholar]
230. Shimazaki T, Shingo T, Weiss S. The ciliary neurotrophic factor/leukemia inhibitory factor/gp130 receptor complex operates in the maintenance of mammalian forebrain neural stem cells. J Neurosci 21: 7642–7653, 2001 [Abstract] [Google Scholar]
231. Shinoda K, Lei H, Yoshii H, Nomura M, Nagano M, Shiba H, Sasaki H, Osawa Y, Ninomiya Y, Niwa O, Morohashi K-I, Li E. Developmental defects of the ventromedial hypothalamic nucleus and pituitary gonadotroph in the Ftz-F1 disrupted mice. Dev Dyn 204: 22–29, 1995 [Abstract] [Google Scholar]
232. Silhol M, Bonnichon V, Rage F, Tapia-Arancibia L. Age-related changes in brain-derived neurotrophic factor and tyrosine kinase receptor isoforms in the hippocampus and hypothalamus in male rats. Neuroscience 132: 613–624, 2005 [Abstract] [Google Scholar]
233. Smith MA, Makino S, Kim SY, Kvetnansky R. Stress increases brain-derived neurotropic factor messenger ribonucleic acid in the hypothalamus and pituitary. Endocrinology 136: 3743–3750, 1995 [Abstract] [Google Scholar]
234. Sobreviela T, Pagcatipunan M, Kroin JS, Mufson EJ. Retrograde transport of brain-derived neurotrophic factor (BDNF) following infusion in neo- and limbic cortex in rat: relationship to BDNF mRNA expressing neurons. J Comp Neurol 375: 417–444, 1996 [Abstract] [Google Scholar]
235. Speliotes EK, Willer CJ, Berndt SI, Monda KL, Thorleifsson G, Jackson AU, Allen HL, Lindgren CM, Luan J, Magi R, Randall JC, Vedantam S, Winkler TW, Qi L, Workalemahu T, Heid IM, Steinthorsdottir V, Stringham HM, Weedon MN, Wheeler E, Wood AR, Ferreira T, Weyant RJ, Segre AV, Estrada K, Liang L, Nemesh J, Park JH, Gustafsson S, Kilpelainen TO, Yang J, Bouatia-Naji N, Esko T, Feitosa MF, Kutalik Z, Mangino M, Raychaudhuri S, Scherag A, Smith AV, Welch R, Zhao JH, Aben KK, Absher DM, Amin N, Dixon AL, Fisher E, Glazer NL, Goddard ME, Heard-Costa NL, Hoesel V, Hottenga JJ, Johansson A, Johnson T, Ketkar S, Lamina C, Li S, Moffatt MF, Myers RH, Narisu N, Perry JR, Peters MJ, Preuss M, Ripatti S, Rivadeneira F, Sandholt C, Scott LJ, Timpson NJ, Tyrer JP, van Wingerden S, Watanabe RM, White CC, Wiklund F, Barlassina C, Chasman DI, Cooper MN, Jansson JO, Lawrence RW, Pellikka N, Prokopenko I, Shi J, Thiering E, Alavere H, Alibrandi MT, Almgren P, Arnold AM, Aspelund T, Atwood LD, Balkau B, Balmforth AJ, Bennett AJ, Ben-Shlomo Y, Bergman RN, Bergmann S, Biebermann H, Blakemore AI, Boes T, Bonnycastle LL, Bornstein SR, Brown MJ, Buchanan TA, Busonero F, Campbell H, Cappuccio FP, Cavalcanti-Proença C, Chen Y-D I, Chen C-M, Chines PS, Clarke R, Coin L, Connell J, Day INM, den Heijer M, Duan J, Ebrahim S, Elliott P, Elosua R, Eiriksdottir G, Erdos MR, Eriksson JG, Facheris MF, Felix SB, Fischer-Posovszky P, Folsom AR, Friedrich N, Freimer NB, Fu M, Gaget S, Gejman PV, Geus EJC, Gieger C, Gjesing AP, Goel A, Goyette P, Grallert H, Gräβler J, Greenawalt DM, Groves CJ, Gudnason V, Guiducci C, Hartikainen A-L, Hassanali N, Hall AS, Havulinna AS, Hayward C, Heath AC, Hengstenberg C, Hicks AA, Hinney A, Hofman A, Homuth G, Hui J, Igl W, Iribarren C, Isomaa B, Jacobs KB, Jarick I, Jewell E, Ulrich J, Jørgensen T, Jousilahti P, Jula A, Kaakinen M, Kajantie E, Kaplan LM, Kathiresan S, Kettunen J, Kinnunen L, Knowles JW, Kolcic I, König IR, Koskinen S, Kovacs P, Kuusisto J, Kraft P, Kvaløy K, Laitinen J, Lantieri O, Lanzani C, Launer LJ, Lecoeur C, Lehtimäki T, Lettre G, Liu J, Lokki M-L, Lorentzon M, Luben RN, Ludwig B, MAGIC. Manunta P, Marek D, Marre M, Martin NG, McArdle WL, McCarthy A, McKnight B, Meitinger T, Melander O, Meyre D, Midthjell K, Montgomery GW, Morken MA, Morris AP, Mulic R, Ngwa JS, Nelis M, Neville MJ, Nyholt DR, O'Donnell CJ, O'Rahilly S, Ong KK, Oostra B, Paré G, Parker AL, Perola M, Pichler I, Pietiläinen KH, Platou CGP, Polasek O, Pouta A, Rafelt S, Raitakari O, Rayner NW, Ridderstråle M, Rief W, Ruokonen A, Robertson NR, Rzehak P, Salomaa V, Sanders AR, Sandhu MS, Sanna S, Saramies J, Savolainen MJ, Scherag S, Schipf S, Schreiber S, Schunkert H, Silander K, Sinisalo J, Siscovick DS, Smit JH, Soranzo N, Sovio U, Stephens J, Surakka I, Swift AJ, Tammesoo M-L, Tardif J-C, Teder-Laving M, Teslovich TM, Thompson JR, Thomson B, Tönjes A, Tuomi T, van Meurs JBJ, van Ommen G-J, Vatin V, Viikari J, Visvikis-Siest S, Vitart V, Vogel CIJ, Voight BF, Waite LL, Wallaschofski H, Bragi Walters G, Widen E, Wiegand S, Wild SH, Willemsen G, Witte DR, Witteman JC, Xu J, Zhang Q, Zgaga L, Ziegler A, Zitting P, Beilby JP, Farooqi IS, Hebebrand J, Huikuri HV, James AL, Kähönen M, Levinson DF, Macciardi F, Nieminen MS, Ohlsson C, Palmer LJ, Ridker PM, Stumvoll M, Beckmann JS, Boeing H, Boerwinkle E, Boomsma DI, Caulfield MJ, Chanock SJ, Collins FS, Cupples LA, Smith GD, Erdmann J, Froguel P, Grönberg H, Gyllensten U, Hall P, Hansen T, Harris TB, Hattersley AT, Hayes RB, Heinrich J, Hu FB, Hveem K, Illig T, Jarvelin M-R, Kaprio J, Karpe F, Khaw K-T, Kiemeney LA, Krude H, Laakso M, Lawlor DA, Metspalu A, Munroe PB, Ouwehand WH, Pedersen O, Penninx BW, Peters A, Pramstaller PP, Quertermous T, Reinehr T, Rissanen A, Rudan I, Samani NJ, Schwarz PEH, Shuldiner AR, Spector TD, Tuomilehto J, Uda M, Uitterlinden A, Valle TT, Wabitsch M, Waeber G, Wareham NJ, Watkins H, on behalf of Procardis Consortium. Wilson JF, Wright AF, Zillikens MC, Chatterjee N, McCarroll SA, Purcell S, Schadt EE, Visscher PM, Assimes TL, Borecki IB, Deloukas P, Fox CS, Groop LC, Haritunians T, Hunter DJ, Kaplan RC, Mohlke KL, O'Connell JR, Peltonen L, Schlessinger D, Strachan DP, van Duijn CM, Wichmann H-E, Frayling TM, Thorsteinsdottir U, Abecasis GR, Barroso I, Boehnke M, Stefansson K, North KE, McCarthy MI, Hirschhorn JN, Ingelsson E, Loos RJF. Association analyses of 249,796 individuals reveal 18 new loci associated with body mass index. Nat Genet 42: 937–948, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
236. Squire LR. Memory and the hippocampus: a synthesis from findings with rats, monkeys, and humans. Psychol Rev 99: 195–231, 1992 [Abstract] [Google Scholar]
237. Sternson SM, Shepherd GM, Friedman JM. Topographic mapping of VMH → arcuate nucleus microcircuits and their reorganization by fasting. Nat Neurosci 8: 1356–1363, 2005 [Abstract] [Google Scholar]
238. Stranahan AM, Khalil D, Gould E. Social isolation delays the positive effects of running on adult neurogenesis. Nat Neurosci 9: 526–533, 2006 [Europe PMC free article] [Abstract] [Google Scholar]
239. Stranahan AM, Norman ED, Lee K, Cutler RG, Telljohann RS, Egan JM, Mattson MP. Diet-induced insulin resistance impairs hippocampal synaptic plasticity and cognition in middle-aged rats. Hippocampus 18: 1085–1088, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
240. Tanaka J, Horiike Y, Matsuzaki M, Miyazaki T, Ellis-Davies GC, Kasai H. Protein synthesis and neurotrophin-dependent structural plasticity of single dendritic spines. Science 319: 1683–1687, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
241. Tapia-Arancibia L, Rage F, Givalois L, Arancibia S. Physiology of BDNF: focus on hypothalamic function. Front Neuroendocrinol 25: 77–107, 2004 [Abstract] [Google Scholar]
242. Teillon S, Calderon GA, Rios M. Diminished diet-induced hyperglycemia and dyslipidemia and enhanced expression of PPARalpha and FGF21 in mice with hepatic ablation of brain-derived neurotropic factor. J Endocrinol 205: 37–47, 2010 [Abstract] [Google Scholar]
243. Temple S. The development of neural stem cells. Nature 414: 112–117, 2001 [Abstract] [Google Scholar]
244. Teng HK, Teng KK, Lee R, Wright S, Tevar S, Almeida RD, Kermani P, Torkin R, Chen ZY, Lee FS, Kraemer RT, Nykjaer A, Hempstead BL. ProBDNF induces neuronal apoptosis via activation of a receptor complex of p75NTR and sortilin. J Neurosci 25: 5455–5463, 2005 [Abstract] [Google Scholar]
245. Ter Horst GJ, Luiten PG. Phaseolus vulgaris leuco-agglutinin tracing of intrahypothalamic connections of the lateral, ventromedial, dorsomedial and paraventricular hypothalamic nuclei in the rat. Brain Res Bull 18: 191–203, 1987 [Abstract] [Google Scholar]
246. Thornberry NA, Lazebnik Y. Caspases: enemies within. Science 281: 1312–1316, 1998 [Abstract] [Google Scholar]
247. Timmusk T, Palm K, Metsis M, Reintam T, Paalme V, Saarma M, Persson H. Multiple promoters direct tissue-specific expression of the rat BDNF gene. Neuron 10: 475–489, 1993 [Abstract] [Google Scholar]
248. Tognoli C, Rossi F, Di Cola F, Baj G, Tongiorgi E, Terova G, Saroglia M, Bernardini G, Gornati R. Acute stress alters transcript expression pattern and reduces processing of proBDNF to mature BDNF in Dicentrarchus labrax. BMC Neurosci 11: 4, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
249. Tominaga-Yoshino K, Kondo S, Tamotsu S, Ogura A. Repetitive activation of protein kinase A induces slow and persistent potentiation associated with synaptogenesis in cultured hippocampus. Neurosci Res 44: 357–367, 2002 [Abstract] [Google Scholar]
250. Tominaga-Yoshino K, Urakubo T, Okada M, Matsuda H, Ogura A. Repetitive induction of late-phase LTP produces long-lasting synaptic enhancement accompanied by synaptogenesis in cultured hippocampal slices. Hippocampus 18: 281–293, 2008 [Abstract] [Google Scholar]
251. Toriya M, Maekawa F, Maejima Y, Onaka T, Fujiwara K, Nakagawa T, Nakata M, Yada T. Long-term infusion of brain-derived neurotrophic factor reduces food intake and body weight via a corticotrophin-releasing hormone pathway in the paraventricular nucleus of the hypothalamus. J Neuroendocrinol 22: 987–995, 2010 [Abstract] [Google Scholar]
252. Tozuka Y, Kumon M, Wada E, Onodera M, Mochizuki H, Wada K. Maternal obesity impairs hippocampal BDNF production and spatial learning performance in young mouse offspring. Neurochem Int 57: 235–247, 2010 [Abstract] [Google Scholar]
253. Tozuka Y, Wada E, Wada K. Diet-induced obesity in female mice leads to peroxidized lipid accumulations and impairment of hippocampal neurogenesis during the early life of their offspring. FASEB J 23: 1920–1934, 2009 [Abstract] [Google Scholar]
254. Tran PV, Akana SF, Malkovska I, Dallman MF, Parada LF, Ingraham HA. Diminished hypothalamic bdnf expression and impaired VMH function are associated with reduced SF-1 gene dosage. J Comp Neurol 498: 637–648, 2006 [Abstract] [Google Scholar]
255. Tsao D, Thomsen HK, Chou J, Stratton J, Hagen M, Loo C, Garcia C, Sloane DL, Rosenthal A, Lin JC. TrkB agonists ameliorate obesity and associated metabolic conditions in mice. Endocrinology 149: 1038–1048, 2008 [Abstract] [Google Scholar]
256. Tsuchida A, Nakagawa T, Itakura Y, Ichihara J, Ogawa W, Kasuga M, Taiji M, Noguchi H. The effects of brain-derived neurotrophic factor on insulin signal transduction in the liver of diabetic mice. Diabetologia 44: 555–566, 2001 [Abstract] [Google Scholar]
257. Tsuchida A, Nonomura T, Nakagawa T, Itakura Y, Ono-Kishino M, Yamanaka M, Sugaru E, Taiji M, Noguchi H. Brain-derived neurotrophic factor ameliorates lipid metabolism in diabetic mice. Diabetes Obes Metab 4: 262–269, 2002 [Abstract] [Google Scholar]
258. Tsuchida A, Nonomura T, Ono-Kishino M, Nakagawa T, Taiji M, Noguchi H. Acute effects of brain-derived neurotrophic factor on energy expenditure in obese diabetic mice. Int J Obes Relat Metab Disord 25: 1286–1293, 2001 [Abstract] [Google Scholar]
259. Tyler WJ, Alonso M, Bramham CR, Pozzo-Miller LD. From acquisition to consolidation: on the role of brain-derived neurotrophic factor signaling in hippocampal-dependent learning. Learn Mem 9: 224–237, 2002 [Europe PMC free article] [Abstract] [Google Scholar]
260. Ubieta R, Uribe RM, Gonzalez JA, Garcia-Vazquez A, Perez-Monter C, Perez-Martinez L, Joseph-Bravo P, Charli JL. BDNF up-regulates pre-pro-TRH mRNA expression in the fetal/neonatal paraventricular nucleus of the hypothalamus. Properties of the transduction pathway. Brain Res 1174: 28–38, 2007 [Abstract] [Google Scholar]
261. Ukropec J, Ukropcova B, Kurdiova T, Gasperikova D, Klimes I. Adipose tissue and skeletal muscle plasticity modulates metabolic health. Arch Physiol Biochem 114: 357–368, 2008 [Abstract] [Google Scholar]
262. Unger TJ, Calderon GA, Bradley LC, Sena-Esteves M, Rios M. Selective deletion of Bdnf in the ventromedial and dorsomedial hypothalamus of adult mice results in hyperphagic behavior and obesity. J Neurosci 27: 14265–14274, 2007 [Abstract] [Google Scholar]
263. van Praag H, Schinder AF, Christie BR, Toni N, Palmer TD, Gage FH. Functional neurogenesis in the adult hippocampus. Nature 415: 1030–1034, 2002 [Abstract] [Google Scholar]
264. Verkuyl JM, Hemby SE, Joels M. Chronic stress attenuates GABAergic inhibition and alters gene expression of parvocellular neurons in rat hypothalamus. Eur J Neurosci 20: 1665–1673, 2004 [Abstract] [Google Scholar]
265. Verkuyl JM, Karst H, Joels M. GABAergic transmission in the rat paraventricular nucleus of the hypothalamus is suppressed by corticosterone and stress. Eur J Neurosci 21: 113–121, 2005 [Abstract] [Google Scholar]
266. Wang C, Bomberg E, Billington C, Levine A, Kotz CM. Brain-derived neurotrophic factor in the hypothalamic paraventricular nucleus increases energy expenditure by elevating metabolic rate. Am J Physiol Regul Integr Comp Physiol 293: R992–R1002, 2007 [Abstract] [Google Scholar]
267. Wang C, Bomberg E, Billington C, Levine A, Kotz CM. Brain-derived neurotrophic factor in the hypothalamic paraventricular nucleus reduces energy intake. Am J Physiol Regul Integr Comp Physiol 293: R1003–R1012, 2007 [Abstract] [Google Scholar]
268. Wang C, Bomberg E, Billington CJ, Levine AS, Kotz CM. Brain-derived neurotrophic factor (BDNF) in the hypothalamic ventromedial nucleus increases energy expenditure. Brain Res 1336: 66–77, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
269. Wang C, Bomberg E, Levine A, Billington C, Kotz CM. Brain-derived neurotrophic factor in the ventromedial nucleus of the hypothalamus reduces energy intake. Am J Physiol Regul Integr Comp Physiol 293: R1037–R1045, 2007 [Abstract] [Google Scholar]
270. Wang C, Godar RJ, Billington CJ, Kotz CM. Chronic administration of brain-derived neurotrophic factor in the hypothalamic paraventricular nucleus reverses obesity induced by high-fat diet. Am J Physiol Regul Integr Comp Physiol 298: R1320–R1332, 2010 [Europe PMC free article] [Abstract] [Google Scholar]
271. Wang CB, Kotz C. Involvement of corticotrophin releasing hormone receptor (CRHR) signaling in BDNF-induced reduction of feeding and body weight gain (Abstract). Obesity 16: s104, 2008 [Google Scholar]
272. Wang CG, Godar R, Dai Y, Bainter H, Billington C, Kotz CM. Effect of brain-derived neurotrophic factor (BDNF) in the ventromedial nucleus of the hypothalamus (VMH) on high fat diet-induced obesity. Obesity 2010 Abstract Supplement 18: S78, 2010 [Google Scholar]
273. Wang GJ, Yang J, Volkow ND, Telang F, Ma Y, Zhu W, Wong CT, Tomasi D, Thanos PK, Fowler JS. Gastric stimulation in obese subjects activates the hippocampus and other regions involved in brain reward circuitry. Proc Natl Acad Sci USA 103: 15641–15645, 2006 [Abstract] [Google Scholar]
274. Westerman MA, Cooper-Blacketer D, Mariash A, Kotilinek L, Kawarabayashi T, Younkin LH, Carlson GA, Younkin SG, Ashe KH. The relationship between Abeta and memory in the Tg2576 mouse model of Alzheimer's disease. J Neurosci 22: 1858–1867, 2002 [Abstract] [Google Scholar]
275. Winocur G, Greenwood CE. The effects of high-fat diets and environmental influences on cognitive performance in rats. Behav Brain Res 101: 153–161, 1999 [Abstract] [Google Scholar]
276. Winocur G, Greenwood CE. Studies of the effects of high fat diets on cognitive function in a rat model. Neurobiol Aging 26 Suppl 1: 46–49, 2005 [Abstract] [Google Scholar]
277. Woo NH, Teng HK, Siao CJ, Chiaruttini C, Pang PT, Milner TA, Hempstead BL, Lu B. Activation of p75NTR by proBDNF facilitates hippocampal long-term depression. Nat Neurosci 8: 1069–1077, 2005 [Abstract] [Google Scholar]
278. Wu A, Molteni R, Ying Z, Gomez-Pinilla F. A saturated-fat diet aggravates the outcome of traumatic brain injury on hippocampal plasticity and cognitive function by reducing brain-derived neurotrophic factor. Neuroscience 119: 365–375, 2003 [Abstract] [Google Scholar]
279. Wu A, Ying Z, Gomez-Pinilla F. The interplay between oxidative stress and brain-derived neurotrophic factor modulates the outcome of a saturated fat diet on synaptic plasticity and cognition. Eur J Neurosci 19: 1699–1707, 2004 [Abstract] [Google Scholar]
280. Wu B, Hu S, Yang M, Pan H, Zhu S. CART peptide promotes the survival of hippocampal neurons by upregulating brain-derived neurotrophic factor. Biochem Biophys Res Commun 347: 656–661, 2006 [Abstract] [Google Scholar]
281. Wu X, Gao J, Yan J, Owyang C, Li Y. Hypothalamus-brain stem circuitry responsible for vagal efferent signaling to the pancreas evoked by hypoglycemia in rat. J Neurophysiol 91: 1734–1747, 2004 [Abstract] [Google Scholar]
282. Wyllie AH, Kerr JF, Currie AR. Cell death: the significance of apoptosis. Int Rev Cytol 68: 251–306, 1980 [Abstract] [Google Scholar]
283. Xu B, Gottschalk W, Chow A, Wilson RI, Schnell E, Zang K, Wang D, Nicoll RA, Lu B, Reichardt LF. The role of brain-derived neurotrophic factor receptors in the mature hippocampus: modulation of long-term potentiation through a presynaptic mechanism involving TrkB. J Neurosci 20: 6888–6897, 2000 [Europe PMC free article] [Abstract] [Google Scholar]
284. Xu B, Goulding EH, Zang K, Cepoi D, Cone RD, Jones KR, Tecott LH, Reichardt LF. Brain-derived neurotrophic factor regulates energy balance downstream of melanocortin-4 receptor. Nat Neurosci 6: 736–742, 2003 [Europe PMC free article] [Abstract] [Google Scholar]
285. Xu J, Stanislaus S, Chinookoswong N, Lau YY, Hager T, Patel J, Ge H, Weiszmann J, Lu SC, Graham M, Busby J, Hecht R, Li YS, Li Y, Lindberg RA, Veniant MM. Acute glucose-lowering and insulin-sensitizing action of FGF21 in insulin resistant mouse models. Association with liver and adipose tissue effects. Am J Physiol Endocrinol Metab 297: E1105–E1114, 2009 [Abstract] [Google Scholar]
286. Yamanaka M, Itakura Y, Inoue T, Tsuchida A, Nakagawa T, Noguchi H, Taiji M. Protective effect of brain-derived neurotrophic factor on pancreatic islets in obese diabetic mice. Metabolism 55: 1286–1292, 2006 [Abstract] [Google Scholar]
287. Yamanaka M, Itakura Y, Ono-Kishino M, Tsuchida A, Nakagawa T, Taiji M. Intermittent administration of brain-derived neurotrophic factor (BDNF) ameliorates glucose metabolism and prevents pancreatic exhaustion in diabetic mice. J Biosci Bioeng 105: 395–402, 2008 [Abstract] [Google Scholar]
288. Yamanaka M, Itakura Y, Tsuchida A, Nakagawa T, Noguchi H, Taiji M. Comparison of the antidiabetic effects of brain-derived neurotrophic factor and thiazolidinediones in obese diabetic mice. Diabetes Obes Metab 9: 879–888, 2007 [Abstract] [Google Scholar]
289. Yamanaka M, Itakura Y, Tsuchida A, Nakagawa T, Taiji M. Brain-derived neurotrophic factor (BDNF) prevents the development of diabetes in prediabetic mice. Biomed Res (Tokyo) 29: 147–153, 2008 [Abstract] [Google Scholar]
290. Yamanaka M, Tsuchida A, Nakagawa T, Nonomura T, Ono-Kishino M, Sugaru E, Noguchi H, Taiji M. Brain-derived neurotrophic factor enhances glucose utilization in peripheral tissues of diabetic mice. Diabetes Obes Metab 9: 59–64, 2007 [Abstract] [Google Scholar]
291. Yeo GS, Connie Hung CC, Rochford J, Keogh J, Gray J, Sivaramakrishnan S, O'Rahilly S, Farooqi IS. A de novo mutation affecting human TrkB associated with severe obesity and developmental delay. Nat Neurosci 7: 1187–1189, 2004 [Abstract] [Google Scholar]
292. Yu Y, Wang Q, Huang XF. Energy-restricted pair-feeding normalizes low levels of brain-derived neurotrophic factor/tyrosine kinase B mRNA expression in the hippocampus, but not ventromedial hypothalamic nucleus, in diet-induced obese mice. Neuroscience 160: 295–306, 2009 [Abstract] [Google Scholar]
293. Zakharenko SS, Patterson SL, Dragatsis I, Zeitlin SO, Siegelbaum SA, Kandel ER, Morozov A. Presynaptic BDNF required for a presynaptic but not postsynaptic component of LTP at hippocampal CA1-CA3 synapses. Neuron 39: 975–990, 2003 [Abstract] [Google Scholar]

Articles from American Journal of Physiology - Regulatory, Integrative and Comparative Physiology are provided here courtesy of American Physiological Society

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/1296698
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/1296698

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1152/ajpregu.00776.2010

Supporting
Mentioning
Contrasting
6
160
0

Article citations


Go to all (149) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.

Funding 


Funders who supported this work.

NIDDK NIH HHS (2)