Open access peer-reviewed chapter

Diet-Induced Overweight Conditions: Effect on Brain Structure, Cognitive Function, and Neurogenesis

Written By

Amina Khatun, Surendra Patra, Kuntal Ghosh, Shrabani Pradhan and Sudipta Chakrabarti

Submitted: 22 November 2022 Reviewed: 20 February 2023 Published: 31 March 2023

DOI: 10.5772/intechopen.110610

From the Edited Volume

Obesity - Recent Insights and Therapeutic Options

Edited by Samy I. McFarlane

Chapter metrics overview

172 Chapter Downloads

View Full Metrics

Abstract

Obesity, a chronic condition that is currently prevalent in both developed and developing nations, is associated with pathological features that ultimately put individuals at risk for a number of negative health issues. Cognitive decline and insulin resistance are two aspects of metabolic syndrome that are closely linked to neurological dysfunction during obesity. Several studies suggest that obesity is associated with regional structural changes, especially signs of cortical thinning in specific brain regions like the hippocampus, and reduced microstructural integrity of the white matter tract is associated with an overall lower academic performance. Obesity causes a loss of brain size and volume indicating a loss of neurons which leads to poor cognitive performance and reduced neurogenesis. An increase in the production of free fatty acids seen with HFD eating might result in increased oxidative stress and increased production of reactive oxygen species. The main cause of systemic inflammation in obesity is the build-up of adipose as it releases TNFα, PAI-1, CRP, IL-1β, and IL-6 which contribute to a pro-inflammatory state in the central nervous system. These elements can all lead to the central IKK/NF-B inflammatory signalling cascade being activated, which can cause a vicious inflammatory cycle that quickens and causes neurodegeneration and cognitive decline.

Keywords

  • obesity
  • oxidative stress
  • inflammation
  • neurodegeneration
  • cognitive loss

1. Introduction

Overweight/obesity, a disease condition currently reaching epidemic proportions, particularly in industrialised countries, and linked to pathological changes that ultimately put people at risk for a variety of adverse health effects [1]. Obesity results from the accumulation of excessive body fat due to the consumption of excessive calories and is typically fuelled by western food habits and a sedentary lifestyle [2]. Food that has been processed and refined typically contains saturated fats, added sugar, and salts, which over long-term contribute to increased calorie intake. There is much speculation and considerable debate about whether obesity-associated cognitive function is a result of weight gain, or whether it is a result of behaviours that lead to weight gain (such as hedonic overeating) [3, 4]. Body mass index (BMI) 25 and 30 are the World Health Organisation’s (WHO) definitions of overweight and obesity, respectively [5]. Body mass index (BMI, kg/m2), which divides weight (kg) by height squared (m2), is widely used to determine obesity. Based on BMI, there are three categories: normal weight (BMI 18.5–24.9 kg/m2), overweight (BMI 25.0–29.9 kg/m2), and obese (BMI 30.0 kg/m2).

However, metabolic syndromes arise when caloric intake significantly outweighs expenditure while there is a sustained lack of physical activity. The presence of low-grade metabolic inflammation in visceral adipose tissue is also considered to be a primary component of metabolic syndrome, which is characterised by central obesity [6]. The metabolic syndrome consists of a number of risk factors that contribute to chronic non-communicable disorders such as cardiovascular diseases (CVD), type 2 diabetes, dyslipidaemia and hypertension as well as other diseases. In addition to CVD, obesity has also been linked to pathological changes in brain morphology and function and cognitive impairments [7]. Even though the central nervous system (CNS) and the peripheral nervous system (PNS) have very different structures and functions, both are vulnerable to the deleterious effects of obesity, indicating that visceral adiposity may facilitate a common pathophysiologic mechanism. The metabolic syndrome includes components that are strongly associated with neurological dysfunction, such as insulin resistance and hypertension leading to cognitive decline. Such factors are generated by obesity. As a result, several processes may be working together to cause neurological dysfunction, while it can be difficult to determine the precise impact of visceral adiposity on neurological dysfunction. Animal obesity models’ mechanistic insights reveal that excessive dietary fat impairs the hypothalamic coordination of energy homeostasis [8]. This pathway may be associated with the disintegration of adipose tissue, resulting in increased levels of free fatty acids, systemic inflammation and dyslipidaemia. As a consequence of chronic calorie overconsumption, circulating triglycerides increase and are adversely affected by numerous organs, including the liver. Dyslipidaemia caused by free fatty acids can result in neurological dysfunction because of lipotoxicity and altered intracellular signalling. Despite these mechanisms affecting the CNS and PNS in multiple ways, obesity is known to have neurological complications [9].

While wide range of functions in human brain requires significant energy, there are limited energy stores in the brain, and those stores can meet only a portion of its energy requirements. The blood-brain barrier (BBB) allows nutrients to continuously enter the brain from the blood, which is how the brain obtains the necessary energy for optimal functioning. Under physiological circumstances, the brain’s primary fuel source is glucose [10]. During development, and at times when glucose supply is insufficient, the brain can use alternative energy substrates such as ketone bodies, triglycerides, as well as lactate [11]. Several recent studies have demonstrated that nutrition and dietary intake have a profound effect on cognition, neuronal function, neuronal signalling, and synaptic plasticity [12, 13]. The consumption of diets rich in saturated fats has consistently been linked to impaired cognitive function, both in clinical and preclinical studies [14, 15]. It is of interest to note that adult studies suggest that consuming high-fat diets (HFD) on a short-term basis may impair attention and memory capabilities [16, 17]. Remarkably, consumption of a diet high in saturated fats is frequently cited as a cause of the cognitive decline and the development of Alzheimer’s Disease [18]. Conversely, it has been suggested that diets high in polyunsaturated fatty acids enhance cognitive health Disease [18, 19, 20]. Depending on the composition, abundance, or lack of particular nutrients, the brain can be affected in different ways by different nutrient requirements. There is a known association between consuming saturated fat-rich diets and metabolic and cardiovascular diseases [21, 22, 23]. According to a growing body of research, obese individuals, and those with diabetes and with hypertension are at high risk of developing cognitive impairments and Alzheimer’s disease [18, 24, 25, 26, 27]. Since the global burden of both neurological diseases and metabolic disorders is increasing, it is necessary to investigate common underlying mechanisms associated with these rapidly increasing disease entities. In this article, we discuss how diet-induced overweight/obesity affect the function, neurogenesis and brain composure.

Advertisement

2. Overweight/obesity: the pathophysiology

In addition to storing energy and releasing it, adipose tissue serves as an insulation system for internal organs and a protection against trauma. Adipokines mediate the endocrine function of adipose tissue through hormones, cytokines, acute phase reactants, and growth factors [28]. These molecules play an important role in maintaining energy homeostasis along with the liver, pancreas, and brain. The major mechanism for achieving energy balance is via controlling energy intake and energy expenditure. Calories are truly calories, and they are all equal according to this fundamental energetic equation [29]. However, when we consider the pathophysiology of obesity-related comorbidities in addition to this merely energy balance aspect, we see that not all calories are created equal [29]. In order to properly explain the pathophysiology of obesity, two simultaneous discussions—one from an energy perspective and the other from a nutritional perspective—must be included. Here, we primarily concentrate on the second because there is controversy regarding the optimal nutritional composition, whereas there is significant agreement on the principles of energy balance management [6, 30]. Managing obesity-related diseases, such as CVD, requires a clear understanding of obesity-independent and obesity-dependent pathophysiologic effects.

Based on genome-wide association studies (GWAS), more than 140 chromosomal regions are associated with obesity [31]. The central nervous system has a significantly enriched gene expression profile associated with BMI and overall obesity [32]. Nevertheless, only a small number of genes, have been found to have a significant impact on BMI thus far. These are the paternally expressed genes along a specific region of chromosome 15 that cause Prader-Willi syndrome, as well as the genes that encode elements of leptin and melanocortin signalling [33]. Most researchers concur that environment, lifestyle, and genetic predisposition contribute to obesity predisposition [34]. Scientists generally agree that increased body weight or adiposity is actively regulated and mitigated by the body under constant environmental conditions, regardless of short-term perturbations in weight or adiposity [35]. Researchers have found that obesity is often defended as a disease, diverting blame from the person to the body’s physiology, just as it is for normal-weight subjects [36]. In addition to adipocytes, adipose tissue contains stromovascular compartments, which are made up of nerve endings, blood vessels, preadipocytes, fibroblasts, endothelial cells, and resident immune cells [37]. As fat-storing cells, adipocytes store triglyceride-rich lipid droplets as a source of energy. Insulin is primarily responsible for regulating adipocyte energy uptake and storage, as it mediates fatty acid influx and lipogenesis while inhibiting lipolysis [38]. When adipocytes are in a negative energy balance, sympathetic neural stimulation promotes lipolysis of stored triglycerides where fatty acids are delivered into the bloodstream to feed non-adipose tissues [39]. The adipocyte undergoes hypertrophy (enlargement of adipocytes) or hyperplasia (proliferation and differentiation) when there is a positive energy balance (i.e., excess caloric intake), it causes the adipocytes to expand in order to store excess calories.

Low-grade metabolic inflammation is associated with increased adipose mass, especially visceral depots, which promote metabolic disease as a consequence of malfunctioning adipose tissue. By increasing adipose mass, especially in visceral depots, metabolic inflammation is linked to metabolic disease, which occurs when adipose tissues malfunction [6]. Adipose tissue inflammation has negative effects on adipokine release, insulin signalling, triglyceride accumulation, and basal lipolysis, among other things. These alterations produce peripheral-tissue and nervous system dysfunction because they result in elevated levels of circulating adipokines and free fatty acids Smith [40]. A chronic caloric surplus that triggers stress signalling pathways and activates local macrophages causes metabolic inflammation, which mostly affects hypertrophied adipose tissue resulting overweight and obesity (Figure 1).

Figure 1.

Pathophysiology of obesity.

Advertisement

3. The relationship between obesity and structural and functional changes in the brain

Various structural changes in the brain can be measured by changes in brain volume or density. Several studies suggest that obesity is associated with regional structural changes, especially in elderly populations [41]. Researchers have reported reduced frontal lobe, anterior cingulate gyrus, hippocampus, and thalamus volume in cognitively healthy obese older individuals. Middle-aged adults and the elderly with high BMI have also been found to have impaired frontal lobe integrity [42, 43, 44]. The volume and density of the brain are often used as indicators of structural changes. A growing body of evidence indicates obesity is associated with regional structural changes in obese populations, especially in elderly people [41]. The hippocampus, cingulate gyrus, and frontal lobes of obese older individuals were found to have reduced volume according to a tensor-based morphometry study [44]. Middle-aged adults and the elderly with high BMI have also been shown to have damage to their frontal lobes [42, 43]. The presence of obesity has been also linked to global structural changes in the brain, including an overall reduction in grey matter and white matter volumes [44].

Grey matter volume structural anomalies in obese patients were discovered by a recent systematic review [45]. The left middle frontal gyrus left middle temporal gyrus, left amygdala, and left cerebellar hemisphere all showed a consistent decline in grey matter in obese people when compared to the control regions, according to an analysis of 10 research published up to December 2017 [41, 43, 46, 47, 48, 49, 50, 51, 52, 53]. A study by Kurth et al. found that the superior frontal gyrus on the left, middle and inferior frontal gyri, the right frontal pole, the left insula, as well as the bilateral superior and middle temporal gyri were negatively affected by body mass index [54]. According to García García et al., obesity and body mass are associated with significantly less grey matter volume in the areas of the brain that play a crucial role in executive control [55]. Obesity-related factors are consistently linked to decreased grey matter volume in a number of regions, including the left temporal pole, bilateral cerebellum, and medial prefrontal cortex. Similar to lean and overweight persons with increasing BMI, obese people have less total grey matter volume. Yokum et al., found that future weight increase is associated with a reduced amount of grey matter in the areas involved in inhibitory regulation [56]. Weight gain is primarily the result of abnormalities in the white matter volumes of the regional spine, not in the grey matter volumes, whereas abnormalities in grey matter volumes increase the likelihood of weight gain in the future.

Similarly, here is compelling evidence that people with increased BMI experience a brain-wide white matter decrease [57, 58]. The result is in line with a major investigation that found links between increased BMI and decreased white matter integrity in two separate, sizable populations [59]. Uncinate fascicle, internal capsule, corticospinal tract, inferior fronto-occipital fascicle, inferior and superior longitudinal fascicles, corpus callosum (cingulate gyrus and hippocampus), and cingulum are just a few of the white matter regions that are known to decrease with a higher BMI [45, 48, 50]. The critical limbic structures are connected to the prefrontal regions by local changes in the white matter fibre tracts linked to greater BMI, which may help to explain why obesity in older age is associated with an increased risk for cognitive impairments and dementia [60]. Obese people could age more quickly than average people, which is thought to raise the risk of cognitive impairment [61]. The fibre tracts that link limbic systems to prefrontal regions are the most commonly affected. These abnormalities are indicative of a loss of white matter integrity brought on by demyelination or inflammatory effects, and can be described by axonal injury or cellular death [61]. The bilateral thalamus, putamen, and globus pallidus in obese individuals are larger than those of normal weight, although the bilateral caudate is smaller [62]. Even obese people (with a BMI of 25 to 30 kg/m2) have symptoms of basal ganglia atrophy and radiating crown [44].

Yau et al. [63] found evidence of cortical thinning in some areas of the brain and decreased microstructural quality of the white matter circuit in obese adolescents. In this study, obese youths performed no worse in cognitive tests than non-obese youths, but structural impairments were associated with a lower academic performance overall [63]. While these findings, do not prove causality, it is biologically conceivable to connect brain anatomical alterations to impaired cognitive function. Poor cognitive performance can be linked causally to smaller brain sizes and volumes since these changes are suggestive of a loss of neurons. Adolescents who are obese may show signs of brain abnormalities, but it may take them until later in life for these changes to cause cognitive impairment. In the early phases of cognitive decline, advanced brain imaging techniques that are more likely to detect anatomical micro alterations do not immediately translate into cognitive dysfunction [64]. As a result, it is particularly important to examine how obesity affects brain function by integrating physiological assessment with cognitive tests. Van Opstal [65] investigated the impact of weight loss (sustained fasting) on brain function in obese persons [65]. Fourteen subjects in this study met the BMI criteria for being classified as obese. During an overnight fast, a 48-hour fast, and an 8-week weight loss programme, brain imaging data were collected using whole-brain resting-state functional magnetic resonance imaging (MRI). The weight loss intervention, according to the researchers, decreased activation in the brain regions in charge of salience, sensory-motor control, and executive control, indicating a connection between weight loss and changes in neurological activity brought on by obesity [65].

A recent meta-analysis on anomalies in brain structure and obesity was undertaken by Opel et al. who also took into account the effects of ageing, hereditary risk, and psychiatric problems [66]. This study involved 6420 participants and examined the relationship between obesity (BMI > 30 kg/m2) and brain anatomy. Results showed a strong relationship between obesity and cortical and subcortical abnormalities, particularly in the lower temporal-frontal cortex thickness. Cortical thickness was influenced by the combination of age and a higher polygenic risk score [66].

Advertisement

4. Cognitive effects of diet-induced obesity/overweight

Cognitive impairment can be caused by obesity when the physiology of the human energy system is impaired. In addition to affecting cognition and the Central Nervous System (CNS), obesity may also affect verbal learning, executive function, and decision-making [67]. An individual’s cognitive function plays an important role in acquiring knowledge and information through the constant use of language, memory, and attention [68]. The effects of obesity on cognitive function can be attributed to structural and functional changes in the brain [69, 70, 71]. In executive functioning tests, obese females performed worse than normal-weight females. There was a reduction in grey matter volume in the left orbitofrontal region associated with a decrease in executive functioning [71]. Older people who are obese during midlife are more likely to develop dementia [72]. Several CVD risks factors, including obesity, T2D, dyslipidaemia, and hypertension, were demonstrated to negatively impact cognition in a systematic review [73]. According to studies, older women who have more body fat have poorer cognitive functioning [71].

Advertisement

5. Childhood obesity and cognitive functions

With childhood obesity is currently on the rise, deficits in attention and cognitive flexibility have been linked to childhood obesity [74]. According to cross-sectional research on overweight children, overweight status was linked to worse test scores, particularly in the areas of arithmetic, reading, and executive function, while being physically fit was linked to better cognition performance, and behaviour [75]. Children who are overweight struggle with spatial cognitive tasks, and studies have revealed variations in both motor and mental rotation efficiency [76]. When rotation activities were challenging, overweight children made more mistakes than children of average weight [76]. Obesity can affect cognitive abilities, particularly executive abilities, in children and adolescents [77, 78]. According to studies, adolescents who are obese have lower executive and attentional cognitive abilities [79, 80]. Insufficient cognitive domains, including executive function and attention deficits, were found in obese adolescents in pilot research [79].

Data analysis from the general population revealed links between gender-specific features of developmental functioning with obesity and impairment [81]. Infants with high subcutaneous fat and children who are overweight or obese had delayed motor development [82]. It has been demonstrated that overweight children have significantly worse perceived and actual physical competence [83]. Overweight children had varying levels of difficulty with basic motor skills [84]. Compared to their peers who were of a healthy weight, obese children reported reduced degrees of gross motor function. The most significant variations were found in balance and locomotor abilities [85]. In comparison to children who were of a healthy weight, children who were overweight performed less effectively in the domains of intelligence, coordination, and gross motor abilities [86]. Significant abnormalities in gross and fine motor skills related to weight were observed in obese children [87].

5.1 Possible mechanisms for cognitive impairment driven by diet-induced overweight

5.1.1 Role of oxidative stress on cognitive impairment

The brain is thought to be the organ most susceptible to harm from oxidative stress [88]. This is explained by the greater lipid content, the high oxygen demand, and the scarcity of antioxidant enzymes [11]. Consistent overnutrition brought on by HFD diet may result in an abundance of reactive oxygen species (ROS) and reactive nitrogen species (RNS) [89]. Because of the excessive production of these species, macromolecules like DNA, membrane lipids, and protein structures are harmed [90]. Reduced glutathione, catalase, and superoxide dismutase act as endogenous antioxidant enzymes to neutralise these free radicals. A high fat diet causes excessive ROS production that exceeds the antioxidant enzymes’ capacity [91].

Here, it is briefly mentioned how an HFD diet might leads to increased oxidative stress. HFD consumption results in an increase in free fatty acid production. In the normal diet-feeding process, electrons are transferred from cofactors (NADH and FADH2) to complex I of the mitochondrial respiratory chain, where they combine safely with oxygen and protons to form water [92]. Free fatty acids are then oxidised in the mitochondria through the mitochondrial respiratory chain. Superoxide radicals are created when some of these electrons interact with oxygen. Conversely, HFD feeding results in increased mitochondrial -oxidation of FFAs, resulting in increased levels of superoxide anion and an excess electron flow [92, 93]. The natural antioxidant enzymes can also be consumed by ROS, which increases the risk of oxidative injury to the brain [92, 94]. According to a growing body of research, the development of cognitive decline and increased oxidative stress may be correlated, [95, 96, 97, 98]. These findings suggest that increased oxidative stress is a major contributor to the cognitive abnormalities caused by HFD.

Many studies have demonstrated that HFD eating causes the levels of oxidative stress indicators in the brain to increase [99, 100, 101, 102, 103, 104]. One of the most delicate areas of the brain to suffer from selective oxidative stress damage is the neocortex and hippocampal region. The impairment of cognition is also closely related to hippocampal oxidative stress [105, 106, 107, 108]. Researchers have used antioxidant therapy to demonstrate cognitive benefits in support of the idea that oxidative stress may be a potential mechanism underlying the cognitive impairment associated with HFD eating. In HFD-fed mice, deficiencies can be restored by employing this strategy [99, 109, 110].

5.1.2 Role of neuroinflammation on cognitive impairment

Inflammation is controlled in a variety of ways by macrophages, depending on their differentiation level. Traditionally activated macrophages (M1) release proinflammatory cytokines and reactive oxygen species (ROS) to initiate an immune response, whereas alternatively activated macrophages (M2) reduce inflammation, promote tissue remodelling, and release growth factors in the later stages of an immune response [111]. Adipose tissue macrophages in healthy, nonobese humans appear to act similarly to M2 macrophages in that they contain arginase, which limits nitric oxide synthesis and promotes polyamine synthesis, and they release little to no proinflammatory cytokines [112]. Chemokine CCL2, formerly called monocyte chemotactic protein-1 (MCP1), is released by macrophages. TNF and IL-6 are also macrophage-released cytokines [113]. It is possible for M1 and M2 macrophages to coexist, which might result in fibrosis and prolonged inflammation [114]. Systemic inflammation in obesity is the result of build-up of adipose tissue and the increase in the levels of tumour necrosis factor-alpha (TNF-α), plasminogen activator inhibitor-1, C-reactive protein, interleukin-1-beta (IL-1β), and interleukin-6 (IL-6), Adipose tissue, particularly lymphocytes and macrophages, contains hypertrophic adipocytes and immune cells that contribute to inflammation [115, 116]. It is possible that processes of necrotic clearance are similar to inflammatory responses mediated by M1 in obese individuals [117, 118]. Macrophages release chemokines like CC-chemokine ligand 2 (CCL2; formerly known as monocyte chemotactic protein-1 (MCP1)) and cytokines like TNF and IL-6. Type 2 diabetes (T2DM) might result from interference in insulin signalling pathway in adipocytes caused by TNF and IL-6 [119]. Over time, macrophages build up in adipose tissue, and the cytokines they release can cause insulin resistance and T2DM [112, 113]. These inflammatory macrophages can increase atherogenic and CVD risks that are a feature of the metabolic syndrome associated with obesity by overexpressing procoagulant proteins [119].

5.1.3 Effects of gut dysbiosis on obesity-related disorders

Abnormalities in neurochemistry and inflammation may also be caused by the gut microbiota associated with obesity [120, 121]. The modification of the gut microbiota, or dysbiosis, can help to explain obesity since it is a factor that is central to host physiology and environmental stressors (such as diet and lifestyle) [120].

Gut dysbiosis (imbalance in gut microbiota composition caused by host genetics, lifestyle, and exposure to microorganisms) may facilitate diet-induced obesity and metabolic complications through a variety of mechanisms, including immune dysregulation, altered energy regulation, altered gut hormone regulation, and proinflammatory mechanisms (such as lipopolysaccharide endotoxins that cross the gut barrier and enter the portal circulation) [121, 122, 123]. Recent studies have indicated that changes in the composition of the gut and inflammation brought on by a leaky gut may have an impact on the pathophysiology of many disorders, such as depression, chronic fatigue syndrome, obesity, and type 2 diabetes (T2DM) (a loss of intestinal barrier integrity that reduces the ability of the gut to protect the internal environment) [124].

Obesity-related inflammation can impact the amygdala, cerebral cortex, hippocampus, and brain stem [125]. Numerous routes, including alteration of the blood–brain barrier (BBB) and choroid plexuses, have been used to link obesity-related low-grade inflammation to neuroinflammation [126]. Insulin resistance is caused by peripheral inflammation, which is seen in obesity [112, 113]. Although it is widely acknowledged that the brain plays a special role in immunity, there have been some instances of transitions between peripheral and central inflammation. It is also possible to express adipokines in the CNS, where these factors have receptors. Adipose tissue produces adipokines, which are expressed in the CNS as well. Peripherally produced adipokines can influence the CNS by crossing the BBB or changing its physiology by interacting with the cells that make up the BBB [127]. There is a strong correlation between neuroinflammation and oxidative stress and a wide range of chronic neurodegenerative diseases [128]. These processes can be regulated by adipokines. Inflammation in the brain can also result from damage to the BBB with ageing [129]. The most significant contributor to cognitive dysfunction may be neuroinflammation, which might also act as a primary pathogenic mechanism for ageing [130].

There is a connection between the activation states of cytokines and chemokines produced by different cell types in adipose tissue outside the central nervous system (quiescent or activated). Obesity and neurodegeneration are linked via the production of inflammatory cytokines and resistance to insulin-like growth factor 1 (IGF-1) [115, 116, 128, 130, 131]. As a result of central inflammation in obesity, hypothalamic satiety signals are interrupted, which perpetuates overeating and has negative consequences for cognition [132]. Different disorders associated with ageing are also thought to be influenced by chronic inflammation. Peripheral inflammation and associated metabolic abnormalities promote T2DM, insulin resistance, and neurodegenerative diseases [133]. In the abdominal adipose tissue, macrophages promote the synthesis of cytokines and proinflammatory chemokines that can cross the BBB. Interferon-gamma can activate microglia, which then serve as relays for neuroinflammation [134].

A number of pathological mechanisms are exacerbated by hypertension, diabetes, and obesity, including cerebral hypoperfusion and glucose hypometabolism. Neuroinflammation and oxidative-nitrosative stress are triggered by these risk factors. There are several cycles of pathological feedback caused by proinflammatory cytokines, endothelin1, and oxidative-nitrosative stress [135]. These cascades cause neurodegeneration and an increase in neuronal Ca2+ [60]. Long-term damage to mitochondria, proteins, DNA, and fatty acids is promoted by oxidative-nitrosative stress. These elements magnify and sustain a variety of problematic feedback loops [136]. Chronic cerebral hypoperfusion results from dysfunctional energy metabolism (compromised mitochondrial ATP production), formation of -amyloid, endothelial dysfunction, and modification of the BBB [115, 130]. Hypoperfusion deprives the brain of oxygen and nutrition, which are its two most critical trophic factors. As a result, the brain experiences synaptic dysfunction and neuronal death, which causes grey and white matter atrophy [136]. The decline of M2 macrophages in the CNS is associated with many neurodegenerative diseases and the subsequent increase of M1-induced inflammation [134]. Microglia and macrophages express the macrophage-stimulating protein receptor (MST1R). In the periphery, obesity-mediated inflammation is attenuated by activating MST1R with its ligand, a macrophage-stimulating protein. Cleavage to the MST1R ligand in vivo regulates the activation of macrophage-dependent repair (M2) [137]. Apoptosis can be caused by neuroinflammation [138]. Its fundamental physiological function, which helps to maintain homeostasis, is a tightly controlled process of cell death. A variety of proteins, signal transducers, and signalling pathway cascades collaborate to fully implement apoptosis [139]. Numerous disorders’ origin and/or progression are strongly correlated with poor apoptotic regulation [138, 140]. TRAIL, TNF, and Fas ligand (Fas-L) bind to the extracellular domain of DR (transmembrane receptors) to initiate the TNF pathway, the main apoptosis pathway [138, 139, 140]. During an inflammatory response, TNF- and Fas-L can cause certain neurons to apoptose [138].

Caspases, which are cysteine proteases, are activated during apoptosis, which controls all of the morphological changes that distinguish this type of cell death [138]. Activating effector caspases (caspases 3, 6, and 7) via a controlled, irreversible, and self-amplifying proteolytic route begins with activating initiator caspases (caspases 2, 8, or 10) [141].

Advertisement

6. Obesity and impaired neurogenesis

It has now been proven beyond a shadow of a doubt that neurogenesis takes place in specific parts of the adult brain, debunking the long-held belief that brain cells lack any capacity for regeneration [142]. Neuronal stem cells (NSCs) in the CNS of adult mammals may now be isolated and identified toas a result of advances in science and technology. Even though adult neurogenesis was first observed in the 1960s, multiple articles have argued that adult mammalian brains do not show any indications of neurogenesis [143, 144, 145]. Adult hippocampal neurogenesis in mammals, including rodents was not “rediscover[ed]” for another three decades [146, 147]. When these freshly produced cells’ “stem-like” properties were first identified in the 1990s, it was believed that NSCs could auto replicate and give rise to a variety of neural lineages in adult mammalian brains, including neurons, astrocytes, and oligodendrocytes [148, 149, 150]. Later research showed that NSCs were mostly found in the subventricular zone (SVZ) of the lateral ventricles and the subgranular zone (SGZ) of the hippocampus dentate gyrus in the adult central CNS [151]. The functional integrity and plasticity of these brain regions are thought to be maintained by these adult NSCs by initiating neurogenesis in response to intrinsic and extrinsic changes, which is consistent with the observation that neurogenesis was also highly prevalent around these regions in adult mammalian brain [152]. Recent research indicates that multipotent NSCs, which have the capacity to differentiate into new neurons as well as astrocytes and oligodendrocytes, can be reprogrammed to grow in the adult mouse hypothalamus in the medio basal region of the hypothalamus (MBH) and the third ventricle wall [153]. These newly formed neurons in the adult mouse hypothalamus could integrate into the existing neuronal network and have recently been shown to be functionally active [154].

Several decades ago, neurogenesis in the adult brain was considered impossible and was refuted, but it has now been widely acknowledged as a very common phenomenon [146, 148]. Many brain regions are abundant in new neurons and NSCs [146]. The hypothalamus, which houses the body’s neuroendocrine system, has recently been found to be both a rich NSC niche and a hotspot of neurogenesis [153]. It is believed that these hypothalamic NSCs play a key role in the neuroendocrine modulation of whole-body physiology [155, 156, 157] as well as systemic ageing [158]. Through local or extrinsic insults to this hypothalamic region, disruption of their neurogenic function may lead to neurodegenerative manifestations across the neuronal milieu. A specific condition known as neurodegeneration, the loss of cognitive abilities, and hypothalamic stem cell damage, as well as obesity and chronic energy imbalance, have been reported to occur after chronic inflammation or inflammation induced by overnutrition or ageing. Additionally, there is data linking brain diseases to overnutrition. Neuroinflammation is clearly to blame for poor neurogenesis and NSC depletion [159]. Furthermore, chronic inflammation slows neurogenesis, NSC survival, and differentiation, and increases ageing-related decline and neurodegenerative disorders [158, 160, 161, 162, 163] despite providing a necessary defensive mechanism for the body. IKB kinase (IKK) and its downstream nuclear transcription factor NFκB (IKK/NFκB signalling) are part of the proinflammatory axis in the hypothalamus that is exacerbated by overnutrition [60, 164]. It has been found that overeating and age-related activation of the IKK/NFκB signalling pathway all contribute to obesity, chronic energy imbalance, neurotoxicity and cognitive impairment 21, and hypothalamic stem cell degradation [164, 165, 166, 167, 168, 169]. A connection between overnutrition/ageing-induced neuroinflammation and neurodegenerative diseases is further supported by evidence and implications relating to overnutrition-induced neurological diseases including Alzheimer’s (AD) and Parkinson’s (PD) [170, 171, 172, 173].

While neuroinflammation is an important defence mechanism in reaction to infections, illnesses, and brain damage, persistent inflammation compromises the body’s normal defences and promotes the progressive neurological and neurodegenerative illnesses [160, 161, 162, 163]. Additionally affected by neurological conditions and disorders, adult neurogenesis is severely impacted by inflammation [174]. For instance, neuroinflammation has been shown to inhibit neurogenesis in the adult hippocampus; neurogenesis may be restored if inflammation is blocked [159, 175]. Adult mice kept on a protracted high-fat diet (HFD) exhibit markedly reduced neurogenesis in the hypothalamus which may also be due to neuroinflammatory reactions generated by the HFD [176, 177]. Contrarily, short-term HFD intake can result in an increase in hypothalamic neurogenesis, which is most likely an adaptive response of the hypothalamus to offset the detrimental effects of HFD feeding on energy balance [176]. The proliferation and differentiation of htNSCs generated from obese mice were shown to be hindered by Li et al., in their study. Such contortion resulted from the overproduction of inflammatory cytokines such TNF- and IL-1, which are known to potently activate IKK/NFκB and were created as a result of NFκB activation, leading to the activation of an inflammatory axis with a positive feed-forward loop [153]. IKK/NFκB activation significantly reduced in vitro htNSC survival, differentiation, and neurogenesis, while IKK/NFκB pathway inhibition increased htNSC survival, differentiation, and neurogenesis, providing additional and direct evidence of the harmful consequences of inflammation [153].

Even while it is established that hippocampal neurogenesis plays a crucial role in appropriate hippocampus function, learning, and memory, there was some scepticism about the report [178, 179, 180]. Two recent investigations, focusing in particular on the limited population of Pro-opiomelanocortin (POMC) neurons that play essential roles in regulating energy balance, showed that chronic HFD-induced obesity and leptin deficit in mice resulted in a reduction in the adult NSC population and new neuron turnover [153]. Through the activation of multiple pro-inflammatory cascades, including the IKK/NFκB inflammatory axis, chronic HFD eating induces metabolic inflammation in the brain, particularly in the hypothalamus. Li et al. showed that chronic HFD eating in mice resulted in both htNSC depletion and neurogenic dysfunction linked to IKK/NFκB activation. The chronic consequences of metabolic dysfunctions, such as excessive calorie intake, glucose intolerance, insulin resistance, and obesity, were discovered in mice that had been genetically modified to have less NSCs in the MBH (Figure 2) [153].

Figure 2.

Possible mechanism of obesity and cognitive dysfunction.

Advertisement

7. Diet-induced obesity and synaptic plasticity dysfunction

The ability of the nervous system to dynamically modify its function in response to ongoing internal processes or outside events is referred to as the nervous system’s plasticity [181]. It is a normal and important aspect of cognition as well as a key method through which the brain can heal after damage [182]. From fundamental physiological processes to integrated behavioural responses, plasticity can be understood and defined at many distinct levels of function [183, 184]. The physical morphology of synapses is related to structural plasticity [185]. Recent research has indicated that the HFD diet, which contains between 47 and 70% fat, has an impact on the brain’s cognitive function [186, 187, 188, 189, 190, 191, 192, 193, 194]. According to this research, HFD causes negative consequences in various brain regions, including the hippocampus, via activating signalling pathways [195, 196]. HFD may have deleterious effects on memory and mood because it alters the systems that control synaptic transmission and the production of proteins associated with plasticity. HFD also causes obesity, which has an impact on the cellular and molecular mechanisms underlying synaptic plasticity in the brain, which impacts learning, memory, and mood. HFD also impaired brain-derived neurotrophic factors (BDNF), and amyloid precursor protein (APP) in the hippocampus.

7.1 Effects of obesity on BDNF

Neuroinflammation in the brain is a significant contributor to the development of neurodegenerative diseases. BDNF has the ability to regulate it [197]. This protein is required for the proper functional development of brain structures as well as the development and retention of synaptic transmission [198]. In addition to its conventional neurotrophic roles, BDNF also seems to have neuroprotective properties against a number of brain traumas, including as ischemia, traumatic brain injuries, and Alzheimer’s disease [199]. It also significantly contributes to the metabolism of energy by reducing food intake, limiting weight gain, and enhancing locomotion following intracerebroventricular injection [183, 200]. Through molecules like synapsin I and growth-associated protein 43, BDNF can control neural plasticity [201, 202]. Synapsin I promotes BDNF modulation of synaptic vesicle exocytosis of neurotransmitters in addition to stimulating axonal growth and supporting synaptic connections [203]. It has been demonstrated that BDNF activation results in synapsin I phosphorylation [204, 205]. According to studies HFD consumption impairs hippocampus synaptic plasticity and cognitive capacities by regulating BDNF expression [206, 207, 208, 209]. Other research has shown that HFD raises brain oxidative stress, which stimulates neuroinflammation and lowers levels of BDNF [186, 210]. After 4 months of HFD ingestion in mice, decreased levels of the basic synaptic proteins SNAP-25 and post-synaptic density (PSD)-95 may increase the brain’s vulnerability to the negative effects of HFD [186].

7.2 Effects of obesity on APP

The biology of APP may have a role in the association between obesity and cognitive performance [211]. APP can be converted into the two peptides Ab1-40 and Ab1-42 by the brain’s neurons, where it is mostly synthesised [212]. Amyloid plaques, a component of AD, are produced when both types of peptides are together [213]. Recent research suggests that the pathophysiology of obesity may also be influenced by APP expression or function [214]. The expression of APP in adipocyte cell lines and adipose tissue has been documented [215, 216]. More significantly, obese people have elevated plasma levels of adipose APP and Ab1-40 [216, 217]. Studies have shown that greater cholesterol levels induce higher amounts of amyloid- in both AD-transgenic and low-density lipoprotein receptor-deficient animals after 8 weeks of diet-dependent obesity in mice [214, 218, 219]. The APP expression or function modifications may be coordinated between several tissue types [214]. In contrast to visceral and subcutaneous fat, one study demonstrated variations in APP expression in brain cells [193]. Inflammation, macrophage and adipocyte phenotype, and macrophage and adipocyte culture phenotype were examined for comparison with the in vivo changes [193].

Another study discovered that feeding mice a very high-fat diet (HFD) for 5 weeks lowered the amounts of the cytoskeleton-associated protein (Arc), which controls baseline activity, in the cerebral cortex and hippocampus. The latter mice developed brain insulin resistance, and acute insulin stimulation reduced phosphatidylinositol 3-kinase (PI3K)/protein kinase B/p70 ribosomal S6 kinase pathway activity, which in turn reduced activation of Arc protein expression [193].

7.3 Effects of obesity on microglia

In addition to driving the inflammatory response in response to various stimuli, microglial cells also regularly maintain neurotrophic connections by remodelling and optimising synapses [220]. Microglia are said to react to neuroinflammation by releasing several kinds of macrovesicles [221]. For instance, when lipopolysaccharide activates BV2 microglial cells, the pro-inflammatory cytokines tumour necrosis factor and interleukin-6 are released [222]. Immature dendritic spine pattern in CA1 diIlabeled neurons, which showed decreased neurogenic ability and lower levels of the scaffold protein Shank 2, suggest impaired connection following an HFD (60% fat) [223]. The medial prefrontal cortex, a part of the brain crucial for cognitive flexibility, exhibits abnormalities in microglial morphology, synapse loss, and cognitive deficits in early-stage obese rats [189]. Additionally, it enhances synaptosome internalisation and microglial activation [191]. Mice given a 60% HFD were found to have fewer dendritic spines, higher levels of microglial activation, and higher levels of synaptic profiles within the microglia. Additionally, transgenic and pharmaceutical techniques that block microglial activation shield obese animals from cognitive deterioration and dendritic spine loss. Additionally, pharmaceutical reduction of microglia’s phagocytic activity has been found to be adequate to stop cognitive decline [224].

7.4 The role of insulin receptors in memory

A crucial part of controlling body metabolism is insulin. However, insulin modifies neural activity, strengthening synaptic connections and increasing memory function [225, 226, 227]. Insulin is known to have profound effects on neurotransmission [228]. PSD fractions contain insulin receptors, which are heavily concentrated in synaptosomes [229]. The scaffolding proteins shank and PSD-95 may also interact with them through insulin receptor tyrosine kinase substrate IRSp53, which is colocalized with synaptophysin and synapsin 1 [230, 231]. Neurites are promoted by insulin, catecholamine release and uptake is regulated, ligand-gated ion channel trafficking is controlled, gamma-aminobutyric acid, N-methyl-d-aspartic acid (NMDA) and AMPA receptors are expressed and localised, and NMDA and PI3K-Akt are involved in modulating synaptic plasticity [232]. There has been significant evidence that insulin resistance and the metabolic syndrome may cause cognitive impairments through mechanisms such as IPMK-mTOR/Akt, synapto-dendritic molecular neuroanatomy, and spatial working memory [233]. In mice, diets high in fat lead to insulin resistance in the cerebral cortex and hypothalamus. Several animal models have shown that insulin resistance affects cognition-related circuitry and neurotransmission [187, 226].

Studies in animals have found reduced dendritic spines in CA1 as well as impairments in long-term memory associated with HFD feeding [191, 234, 235]. Many different parts of the brain can be negatively impacted by alterations in glucose homeostasis, but the hippocampus is particularly susceptible to them [236, 237, 238]. According to research by Strahan et al., rats fed a high-fat, high-glucose diet for 8 months showed significant impairments in their hippocampal dendritic spine density, spatial learning ability, and LTP at Schaffer collateral-CA1 synapses [239]. In addition, high-fat, high-glucose-fed animals showed a reduction in BDNF levels in the hippocampus compared with controls. They suggested that the changes could be due to peripheral insulin resistance, or that some components of the diet may directly affect brain health and hippocampal plasticity. Further evidence suggests that the DG of rats fed HFD exhibit impaired stimulus-evoked LTP [240]. HFD feeding leads to a change in the expression of protein-coding genes in the cortex, but studies have not explored alterations in non-coding RNAs. The HFD feeding of mice led to changes in both coding and non-coding RNA expression in the brain cortex, according to a study by Yoon and colleagues. According to these researchers, consuming HFD for 8 weeks causes a decrease in the expression of genes linked to synaptogenesis and neurotransmitter release [241]. In the hippocampus of animals receiving HFD, Arnold, and colleagues likewise found that PSD-95 expression was downregulated [187]. Interesting investigations have found that animals given the HFD have less neurogenesis in their dentate gyrus [242]. Thus, it is clear that HFD may significantly modify the expression of many genes linked to these processes, which may have a negative impact on brain morphology and synaptic plasticity.

Advertisement

8. Conclusion

Obesity is a serious public health problem that is increasing an proportions with serious health and societal problems. It mostly affects cognition through changing the structures and operations of the brain. Brain structure, leptin/insulin dysregulation, oxidative stress, cerebrovascular function, blood-brain barrier, and inflammation are all impacted by obesity and contribute to the decline in cognitive performance. Inhibition of neurogenesis is thought to be caused by neuroinflammation, which can be brought on by a variety of internal or external factors, including ER and oxidative stress, as well as, overnutrition-induced metabolic inflammation, and autophagic defects. These factors are all connected to the activation of the central IKK/NF-B inflammatory signalling cascade, which can result in a vicious inflammatory cycle that accelerates ageing, neurodegeneration, and cognitive decline.

References

  1. 1. Mcardle MA, Finucane OM, Connaughton RM, Mcmorrow AM, Roche HM. Mechanisms of obesity-induced inflammation and insulin resistance: Insights into the emerging role of nutritional strategies. Frontiers in Endocrinology. 2013;4:52
  2. 2. Kant AK. Dietary patterns and health outcomes. Journal of the American Dietetic Association. 2004;104:615-635
  3. 3. Mcewen BS, Morrison JH. The brain on stress: Vulnerability and plasticity of the prefrontal cortex over the life course. Neuron. 2013;79:16-29
  4. 4. Schwartz MW, Porte D Jr. Diabetes, obesity, and the brain. Science. 2005;307:375-379
  5. 5. Organization WH. Obesity and overweight. Fact sheet N 311. 2012. Available from: http://www.who.int/mediacentre/factsheets/fs311/en/
  6. 6. Klöting N, Blüher M. Adipocyte dysfunction, inflammation and metabolic syndrome. Reviews in Endocrine and Metabolic Disorders. 2014;15:277-287
  7. 7. Stillman CM, Weinstein AM, Marsland AL, Gianaros PJ, Erickson KI. Body–brain connections: The effects of obesity and behavioral interventions on neurocognitive aging. Frontiers in Aging Neuroscience. 2017;9:115
  8. 8. Timper K, Brüning JC. Hypothalamic circuits regulating appetite and energy homeostasis: Pathways to obesity. Disease Models & Mechanisms. 2017;10:679-689
  9. 9. Do’Brien P, Hinder LM, Callaghan BC, Feldman EL. Neurological consequences of obesity. The Lancet Neurology. 2017;16:465-477
  10. 10. Dienel GA. Fueling and imaging brain activation. ASN Neuro. 2012;4 AN20120021
  11. 11. Bordone MP, Salman MM, Titus HE, Amini E, Andersen JV, Chakraborti B, et al. The energetic brain–a review from students to students. Journal of Neurochemistry. 2019;151:139-165
  12. 12. Gomez-Pinilla F, Tyagi E. Diet and cognition: Interplay between cell metabolism and neuronal plasticity. Current Opinion in Clinical Nutrition and Metabolic Care. 2013;16:726
  13. 13. Mattson MP. Energy intake and exercise as determinants of brain health and vulnerability to injury and disease. Cell Metabolism. 2012;16:706-722
  14. 14. Greenwood CE, Winocur G. High-fat diets, insulin resistance and declining cognitive function. Neurobiology of Aging. 2005;26:42-45
  15. 15. Molteni R, Barnard R, Ying Z, Roberts C, Gomez-Pinilla F. A high-fat, refined sugar diet reduces hippocampal brain-derived neurotrophic factor, neuronal plasticity, and learning. Neuroscience. 2002;112:803-814
  16. 16. Bayer-Carter JL, Green PS, Montine TJ, Vanfossen B, Baker LD, Watson GS, et al. Diet intervention and cerebrospinal fluid biomarkers in amnestic mild cognitive impairment. Archives of Neurology. 2011;68:743-752
  17. 17. Edwards LM, Murray AJ, Holloway CJ, Carter EE, Kemp GJ, Codreanu I, et al. Short-term consumption of a high-fat diet impairs whole-body efficiency and cognitive function in sedentary men. The FASEB Journal. 2011;25:1088-1096
  18. 18. Eskelinen MH, Ngandu T, Helkala EL, Tuomilehto J, Nissinen A, Soininen H, et al. Fat intake at midlife and cognitive impairment later in life: A population-based CAIDE study. International Journal of Geriatric Psychiatry: A Journal of the Psychiatry of Late Life and Allied Sciences. 2008;23:741-747
  19. 19. Mccann JC, Ames BN. Is docosahexaenoic acid, an n− 3 long-chain polyunsaturated fatty acid, required for development of normal brain function? An overview of evidence from cognitive and behavioral tests in humans and animals. The American Journal of Clinical Nutrition. 2005;82:281-295
  20. 20. Wu A, Ying Z, Gomez-Pinilla F. Omega-3 fatty acids supplementation restores mechanisms that maintain brain homeostasis in traumatic brain injury. Journal of Neurotrauma. 2007;24:1587-1595
  21. 21. Crescenzo R, Bianco F, Mazzoli A, Giacco A, Cancelliere R, Di Fabio G, et al. Fat quality influences the obesogenic effect of high fat diets. Nutrients. 2015;7:9475-9491
  22. 22. Kraegen EW, Clark PW, Jenkins AB, Daley EA, Chisholm DJ, Storlien LH. Development of muscle insulin resistance after liver insulin resistance in high-fat–fed rats. Diabetes. 1991;40:1397-1403
  23. 23. Siri-Tarino PW, Sun Q , Hu FB, Krauss RM. Saturated fat, carbohydrate, and cardiovascular disease. The American Journal of Clinical Nutrition. 2010;91:502-509
  24. 24. Biessels GJ, Despa F. Cognitive decline and dementia in diabetes mellitus: Mechanisms and clinical implications. Nature Reviews Endocrinology. 2018;14:591-604
  25. 25. Chatterjee S, Peters SA, Woodward M, Mejia Arango S, Batty GD, Beckett N, et al. Type 2 diabetes as a risk factor for dementia in women compared with men: A pooled analysis of 2.3 million people comprising more than 100,000 cases of dementia. Diabetes Care. 2016;39:300-307
  26. 26. Crane PK, Walker R, Hubbard RA, Li G, Nathan DM, Zheng H, et al. Glucose levels and risk of dementia. New England Journal of Medicine. 2013;369:540-548
  27. 27. de Felice FG, Lourenco MV, Ferreira ST. How does brain insulin resistance develop in Alzheimer’s disease? Alzheimer’s & Dementia. 2014;10:S26-S32
  28. 28. Balistreri CR, Caruso C, Candore G. The role of adipose tissue and adipokines in obesity-related inflammatory diseases. Mediators of Inflammation. 2010;2010
  29. 29. Hill JO, Wyatt HR, Peters JC. Energy balance and obesity. Circulation. 2012;126:126-132
  30. 30. de Souza RJ, Bray GA, Carey VJ, Hall KD, Leboff MS, Loria CM, et al. Effects of 4 weight-loss diets differing in fat, protein, and carbohydrate on fat mass, lean mass, visceral adipose tissue, and hepatic fat: Results from the POUNDS LOST trial. The American Journal of Clinical Nutrition. 2012;95:614-625
  31. 31. Fall T, Mendelson M, Speliotes EK. Recent advances in human genetics and epigenetics of adiposity: Pathway to precision medicine? Gastroenterology. 2017;152:1695-1706
  32. 32. Locke AE, Kahali B, Berndt SI, Justice AE, Pers TH, Day FR, et al. Genetic studies of body mass index yield new insights for obesity biology. Nature. 2015;518:197-206
  33. 33. Angulo M, Butler M, Cataletto M. Prader-Willi syndrome: A review of clinical, genetic, and endocrine findings. Journal of Endocrinological Investigation. 2015;38:1249-1263
  34. 34. Albuquerque D, Nóbrega C, Manco L, Padez C. The contribution of genetics and environment to obesity. British Medical Bulletin. 2017;123:159-173
  35. 35. Schwartz MW, Seeley RJ, Zeltser LM, DRewnowski A, Ravussin E, Redman LM, et al. Obesity pathogenesis: An endocrine society scientific statement. Endocrine Reviews. 2017;38:267-296
  36. 36. Hall KD, Guo J. Obesity energetics: Body weight regulation and the effects of diet composition. Gastroenterology. 2017;152:1718-1727. e3
  37. 37. Rosen ED, Spiegelman BM. What we talk about when we talk about fat. Cell. 2014;156:20-44
  38. 38. Kalupahana NS, Moustaid-Moussa N. The renin-angiotensin system: A link between obesity, inflammation and insulin resistance. Obesity Reviews. 2012;13:136-149
  39. 39. Bartness TJ, Shrestha Y, Vaughan C, Schwartz G, Song C. Sensory and sympathetic nervous system control of white adipose tissue lipolysis. Molecular and Cellular Endocrinology. 2010;318:34-43
  40. 40. Smith MM, Minson CT. Obesity and adipokines: Effects on sympathetic overactivity. The Journal of Physiology. 2012;590:1787-1801
  41. 41. Brooks SJ, Benedict C, Burgos J, Kempton M, Kullberg J, Nordenskjöld R, et al. Late-life obesity is associated with smaller global and regional gray matter volumes: A voxel-based morphometric study. International Journal of Obesity. 2013;37:230-236
  42. 42. Gazdzinski S, Kornak J, Weiner MW, Meyerhoff DJ. Body mass index and magnetic resonance markers of brain integrity in adults. Annals of Neurology. 2008;63:652-657
  43. 43. Pannacciulli N, Del Parigi A, Chen K, Le DSN, Reiman EM, Tataranni PA. Brain abnormalities in human obesity: A voxel-based morphometric study. NeuroImage. 2006;31:1419-1425
  44. 44. Raji CA, Ho AJ, Parikshak NN, Becker JT, Lopez OL, Kuller LH, et al. Brain structure and obesity. Human Brain Mapping. 2010;31:353-364
  45. 45. Herrmann MJ, Tesar AK, Beier J, Berg M, Warrings B. Grey matter alterations in obesity: A meta-analysis of whole-brain studies. Obesity Reviews. 2019;20:464-471
  46. 46. Haltia LT, Viljanen A, Parkkola R, Kemppainen N, Rinne JO, Nuutila P, et al. Brain white matter expansion in human obesity and the recovering effect of dieting. The Journal of Clinical Endocrinology & Metabolism. 2007;92:3278-3284
  47. 47. Jauch-Chara K, Binkofski F, Loebig M, Reetz K, Jahn G, Melchert UH, et al. Blunted brain energy consumption relates to insula atrophy and impaired glucose tolerance in obesity. Diabetes. 2015;64:2082-2091
  48. 48. Karlsson HK, Tuulari JJ, Hirvonen J, Lepomäki V, Parkkola R, Hiltunen J, et al. Obesity is associated with white matter atrophy: A combined diffusion tensor imaging and voxel-based morphometric study. Obesity. 2013;21:2530-2537
  49. 49. Mathar D, Horstmann A, Pleger B, Villringer A, Neumann J. Is it worth the effort? Novel insights into obesity-associated alterations in cost-benefit decision-making. Frontiers in Behavioral Neuroscience. 2016;9:360
  50. 50. Shott ME, Cornier M-A, Mittal VA, Pryor TL, Orr JM, Brown MS, et al. Orbitofrontal cortex volume and brain reward response in obesity. International Journal of Obesity. 2015;39:214-221
  51. 51. Tuulari JJ, Karlsson HK, Antikainen O, Hirvonen J, Pham T, Salminen P, et al. Bariatric surgery induces white and grey matter density recovery in the morbidly obese: A voxel-based morphometric study. Human Brain Mapping. 2016;37:3745-3756
  52. 52. Wang H, Wen B, Cheng J, Li H. Brain structural differences between normal and obese adults and their links with lack of perseverance, negative urgency, and sensation seeking. Scientific Reports. 2017;7:1-7
  53. 53. Zhang B, Tian X, Tian D, Wang J, Wang Q , Yu C, et al. Altered regional gray matter volume in obese men: A structural MRI study. Frontiers in Psychology. 2017;8:125
  54. 54. Kurth F, Levitt JG, Phillips OR, Luders E, Woods RP, Mazziotta JC, et al. Relationships between gray matter, body mass index, and waist circumference in healthy adults. Human Brain Mapping. 2013;34:1737-1746
  55. 55. García-García I, Michaud A, Dadar M, Zeighami Y, Neseliler S, Collins DL, et al. Neuroanatomical differences in obesity: Meta-analytic findings and their validation in an independent dataset. International Journal of Obesity. 2019;43:943-951
  56. 56. Yokum S, Ng J, Stice E. Relation of regional gray and white matter volumes to current BMI and future increases in BMI: A prospective MRI study. International Journal of Obesity. 2012;36:656-664
  57. 57. He Q , Chen C, Dong Q , Xue G, Chen C, Lu Z-L, et al. Gray and white matter structures in the midcingulate cortex region contribute to body mass index in Chinese young adults. Brain Structure and Function. 2015;220:319-329
  58. 58. Papageorgiou I, Astrakas LG, Xydis V, Alexiou GA, Bargiotas P, Tzarouchi L, et al. Abnormalities of brain neural circuits related to obesity: A diffusion tensor imaging study. Magnetic Resonance Imaging. 2017;37:116-121
  59. 59. Repple J, Opel N, Meinert S, Redlich R, Hahn T, Winter NR, et al. Elevated body-mass index is associated with reduced white matter integrity in two large independent cohorts. Psychoneuroendocrinology. 2018;91:179-185
  60. 60. Cai D. Neuroinflammation and neurodegeneration in overnutrition-induced diseases. Trends in Endocrinology & Metabolism. 2013;24:40-47
  61. 61. Kullmann S, Callaghan MF, Heni M, Weiskopf N, Scheffler K, Häring H-U, et al. Specific white matter tissue microstructure changes associated with obesity. NeuroImage. 2016;125:36-44
  62. 62. Kim A-Y, Shim J-H, Choi HJ, Baek H-M. Comparison of volumetric and shape changes of subcortical structures based on 3-dimensional image between obesity and normal-weighted subjects using 3.0 T MRI. Journal of Clinical Neuroscience. 2020;73:280-287
  63. 63. Yau PL, Kang EH, Javier DC, Convit A. Preliminary evidence of cognitive and brain abnormalities in uncomplicated adolescent obesity. Obesity. 2014;22:1865-1871
  64. 64. Ryan JP, Fine DF, Rosano C. Type 2 diabetes and cognitive impairment: Contributions from neuroimaging. Journal of Geriatric Psychiatry and Neurology. 2014;27:47-55
  65. 65. Van Opstal A, Wijngaarden M, Van der Grond J, Pijl H. Changes in brain activity after weight loss. Obesity Science & Practice. 2019;5:459-467
  66. 66. Opel N, Thalamuthu A, Milaneschi Y, Grotegerd D, Flint C, Leenings R, et al. Brain structural abnormalities in obesity: Relation to age, genetic risk, and common psychiatric disorders. Molecular Psychiatry. 2021;26:4839-4852
  67. 67. Wang C, Chan JS, Ren L, Yan JH. Obesity reduces cognitive and motor functions across the lifespan. Neural Plasticity. 2016;2016
  68. 68. Hess LM. Chemotherapy-related change in cognitive function: A conceptual model. 1969;34(5):981-994
  69. 69. Bolzenius JD, Laidlaw DH, Cabeen RP, Conturo TE, Mcmichael AR, Lane EM, et al. Brain structure and cognitive correlates of body mass index in healthy older adults. Behavioural Brain Research. 2015;278:342-347
  70. 70. Volkow ND, Wang GJ, Telang F, Fowler JS, Goldstein RZ, Alia-Klein N, et al. Inverse association between BMI and prefrontal metabolic activity in healthy adults. Obesity. 2009;17:60-65
  71. 71. Walther K, Birdsill AC, Glisky EL, Ryan L. Structural brain differences and cognitive functioning related to body mass index in older females. Human Brain Mapping. 2010;31:1052-1064
  72. 72. Whitmer RA, Gunderson EP, Barrett-Connor E, Quesenberry CP, Yaffe K. Obesity in middle age and future risk of dementia: A 27 year longitudinal population based study. BMJ. 2005;330:1360
  73. 73. Van den Berg E, Kloppenborg RP, Kessels RP, Kappelle LJ, Biessels GJ. Type 2 diabetes mellitus, hypertension, dyslipidemia and obesity: A systematic comparison of their impact on cognition. Biochimica et Biophysica Acta (BBA)-Molecular Basis of Disease. 2009;1792:470-481
  74. 74. Cserjési R, Molnár D, Luminet O, Lénárd L. Is there any relationship between obesity and mental flexibility in children? Appetite. 2007;49:675-678
  75. 75. Davis CL, Tomporowski PD, Mcdowell JE, Austin BP, Miller PH, Yanasak NE, et al. Exercise improves executive function and achievement and alters brain activation in overweight children: A randomized, controlled trial. Health Psychology. 2011;30:91
  76. 76. Jansen P, Schmelter A, Kasten L, Heil M. Impaired mental rotation performance in overweight children. Appetite. 2011;56:766-769
  77. 77. Kamijo K, Khan NA, Pontifex MB, Scudder MR, Drollette ES, Raine LB, et al. The relation of adiposity to cognitive control and scholastic achievement in preadolescent children. Obesity. 2012;20:2406-2411
  78. 78. Verdejo-García A, Pérez-Expósito M, Schmidt-Río-Valle J, Fernández-Serrano MJ, Cruz F, Pérez-García M, et al. Selective alterations within executive functions in adolescents with excess weight. Obesity. 2010;18:1572-1578
  79. 79. Lokken KL, Boeka AG, Austin HM, Gunstad J, Harmon CM. Evidence of executive dysfunction in extremely obese adolescents: A pilot study. Surgery for Obesity and Related Diseases. 2009;5:547-552
  80. 80. Schwartz DH, Leonard G, Perron M, Richer L, Syme C, Veillette S, et al. Visceral fat is associated with lower executive functioning in adolescents. International Journal of Obesity. 2013;37:1336-1343
  81. 81. Mond J, Stich H, Hay P, Krämer A, Baune B. Associations between obesity and developmental functioning in pre-school children: A population-based study. International Journal of Obesity. 2007;31:1068-1073
  82. 82. Slining M, Adair LS, Goldman BD, Borja JB, Bentley M. Infant overweight is associated with delayed motor development. The Journal of Pediatrics. 2010;157:20-25. e1
  83. 83. Southall JE, Okely AD, Steele JR. Actual and perceived physical competence in overweight and nonoverweight children. Pediatric Exercise Science. 2004;16:15-24
  84. 84. Poulsen AA, Desha L, Ziviani J, Griffiths L, Heaslop A, Khan A, et al. Fundamental movement skills and self-concept of children who are overweight. International Journal of Pediatric Obesity. 2011;6:e464-e471
  85. 85. Roberts D, Veneri D, Decker R, Gannotti M. Weight status and gross motor skill in kindergarten children. Pediatric Physical Therapy. 2012;24:353-360
  86. 86. Krombholz H. Motor and cognitive performance of overweight preschool children. Perceptual and Motor Skills. 2013;116:40-57
  87. 87. Gentier I, D’hondt E, Shultz S, Deforche B, Augustijn M, Hoorne S, et al. Fine and gross motor skills differ between healthy-weight and obese children. Research in Developmental Disabilities. 2013;34:4043-4051
  88. 88. Garbarino VR, Orr ME, Rodriguez KA, Buffenstein R. Mechanisms of oxidative stress resistance in the brain: Lessons learned from hypoxia tolerant extremophilic vertebrates. Archives of Biochemistry and Biophysics. 2015;576:8-16
  89. 89. Verdile G, Keane KN, Cruzat VF, Medic S, Sabale M, Rowles J, et al. Inflammation and oxidative stress: The molecular connectivity between insulin resistance, obesity, and Alzheimer’s disease. Mediators of Inflammation. 2015;2015
  90. 90. Tan BL, Norhaizan ME. Effect of high-fat diets on oxidative stress, cellular inflammatory response and cognitive function. Nutrients. 2019;11:2579
  91. 91. Tangvarasittichai S. Oxidative stress, insulin resistance, dyslipidemia and type 2 diabetes mellitus. World Journal of Diabetes. 2015;6:456
  92. 92. Pessayre D, Mansouri A, Fromenty B. V. Mitochondrial dysfunction in steatohepatitis. American Journal of Physiology-Gastrointestinal and Liver Physiology. 2002;282:G193-G199
  93. 93. Matsuzawa-Nagata N, Takamura T, Ando H, Nakamura S, Kurita S, Misu H, et al. Increased oxidative stress precedes the onset of high-fat diet–induced insulin resistance and obesity. Metabolism. 2008;57:1071-1077
  94. 94. Le Lay S, Simard G, Martinez MC, Andriantsitohaina R. Oxidative stress and metabolic pathologies: From an adipocentric point of view. Oxidative Medicine and Cellular Longevity. 2014;2014
  95. 95. Head E. Oxidative damage and cognitive dysfunction: Antioxidant treatments to promote healthy brain aging. Neurochemical Research. 2009;34:670-678
  96. 96. Perry G, Moreira PI, Santos MS, Oliveira CR, Shenk JC, Nunomura A, et al. Alzheimer disease and the role of free radicals in the pathogenesis of the disease. CNS & Neurological Disorders-Drug Targets (Formerly Current Drug Targets-CNS & Neurological Disorders). 2008;7:3-10
  97. 97. Perry G, Nunomura A, Hirai K, Zhu X, Prez M, Avila J, et al. Is oxidative damage the fundamental pathogenic mechanism of Alzheimer’s and other neurodegenerative diseases? Free Radical Biology and Medicine. 2002;33:1475-1479
  98. 98. Su B, Wang X, Nunomura A, Moreira PI, Lee H-G, Perry G, et al. Oxidative stress signaling in Alzheimer’s disease. Current Alzheimer Research. 2008;5:525-532
  99. 99. Kumar A, Jaggi AS, Sodhi RK, Singh N. Silymarin ameliorates memory deficits and neuropathological changes in mouse model of high-fat-diet-induced experimental dementia. Naunyn-Schmiedeberg’s Archives of Pharmacology. 2014;387:777-787
  100. 100. Liu Y, Fu X, Lan N, Li S, Zhang J, Wang S, et al. Luteolin protects against high fat diet-induced cognitive deficits in obesity mice. Behavioural Brain Research. 2014;267:178-188
  101. 101. Lu J, Wu D-M, Zheng Z-H, Zheng Y-L, Hu B, Zhang Z-F. Troxerutin protects against high cholesterol-induced cognitive deficits in mice. Brain. 2011;134:783-797
  102. 102. Pintana H, Apaijai N, Chattipakorn N, Chattipakorn SC. DPP-4 inhibitors improve cognition and brain mitochondrial function of insulin-resistant rats. The Journal of Endocrinology. 2013;218:1-11
  103. 103. Pintana H, Apaijai N, Pratchayasakul W, Chattipakorn N, Chattipakorn SC. Effects of metformin on learning and memory behaviors and brain mitochondrial functions in high fat diet induced insulin resistant rats. Life Sciences. 2012;91:409-414
  104. 104. Sodhi RK, Singh N. Liver X receptor agonist T0901317 reduces neuropathological changes and improves memory in mouse models of experimental dementia. European Journal of Pharmacology. 2014;732:50-59
  105. 105. de Oliveira MR, Rocha RFD, Stertz L, Fries GR, de Oliveira DL, Kapczinski F, et al. Total and mitochondrial nitrosative stress, decreased brain-derived neurotrophic factor (BDNF) levels and glutamate uptake, and evidence of endoplasmic reticulum stress in the hippocampus of vitamin A-treated rats. Neurochemical Research. 2011;36:506-517
  106. 106. Evans JL, Goldfine ID, Maddux BA, Grodsky GM. Are oxidative stress− activated signaling pathways mediators of insulin resistance and β-cell dysfunction? Diabetes. 2003;52:1-8
  107. 107. Farr SA, Yamada KA, Butterfield DA, Abdul HM, Xu L, Miller NE, et al. Obesity and hypertriglyceridemia produce cognitive impairment. Endocrinology. 2008;149:2628-2636
  108. 108. Stranahan AM, Cutler RG, Button C, Telljohann R, Mattson MP. Diet-induced elevations in serum cholesterol are associated with alterations in hippocampal lipid metabolism and increased oxidative stress. Journal of Neurochemistry. 2011;118:611-615
  109. 109. Alzoubi KH, Khabour OF, Salah HA, Hasan Z. Vitamin E prevents high-fat high-carbohydrates diet-induced memory impairment: The role of oxidative stress. Physiology & Behavior. 2013;119:72-78
  110. 110. Xia S-F, Xie Z-X, Qiao Y, Li L-R, Cheng X-R, Duan X-M, et al. Salvianolic acid B counteracts cognitive decline triggered by oxidative stress in mice fed with high-fat diets. Journal of Functional Foods. 2014;11:278-292
  111. 111. O’shea JJ, Murray PJ. Cytokine signaling modules in inflammatory responses. Immunity. 2008;28:477-487
  112. 112. Lumeng CN, Bodzin JL, Saltiel AR. Obesity induces a phenotypic switch in adipose tissue macrophage polarization. The Journal of Clinical Investigation. 2007;117:175-184
  113. 113. Zeyda M, Stulnig TM. Adipose tissue macrophages. Immunology Letters. 2007;112:61-67
  114. 114. Poltavets AS, Vishnyakova PA, Elchaninov AV, Sukhikh GT, Fatkhudinov TK. Macrophage modification strategies for efficient cell therapy. Cell. 2020;9:1535
  115. 115. Hahm JR, Jo MH, Ullah R, Kim MW, Kim MO. Metabolic stress alters antioxidant systems, suppresses the adiponectin receptor 1 and induces Alzheimer’s like pathology in mice brain. Cell. 2020;9:249
  116. 116. Nguyen JC, Killcross AS, Jenkins TA. Obesity and cognitive decline: Role of inflammation and vascular changes. Frontiers in Neuroscience. 2014;8:375
  117. 117. Cinti S, Mitchell G, Barbatelli G, Murano I, Ceresi E, Faloia E, et al. Adipocyte death defines macrophage localization and function in adipose tissue of obese mice and humans. Journal of Lipid Research. 2005;46:2347-2355
  118. 118. Mosser DM, Edwards JP. Exploring the full spectrum of macrophage activation. Nature Reviews Immunology. 2008;8:958-969
  119. 119. Bastard J-P, Maachi M, Lagathu C, Kim MJ, Caron M, Vidal H, et al. Recent advances in the relationship between obesity, inflammation, and insulin resistance. European Cytokine Network. 2006;17:4-12
  120. 120. Agustí A, García-Pardo MP, López-Almela I, Campillo I, Maes M, Romaní-Pérez M, et al. Interplay between the gut-brain axis, obesity and cognitive function. Frontiers in Neuroscience. 2018;12:155
  121. 121. Valdes AM, Walter J, Segal E, Spector TD. Role of the gut microbiota in nutrition and health. BMJ. 2018;361
  122. 122. Baothman OA, Zamzami MA, Taher I, Abubaker J, Abu-Farha M. The role of gut microbiota in the development of obesity and diabetes. Lipids in Health and Disease. 2016;15:1-8
  123. 123. Kim Y-G, Udayanga KGS, Totsuka N, Weinberg JB, Núñez G, Shibuya A. Gut dysbiosis promotes M2 macrophage polarization and allergic airway inflammation via fungi-induced PGE2. Cell Host & Microbe. 2014;15:95-102
  124. 124. Slyepchenko A, Maes M, Jacka FN, Köhler CA, Barichello T, Mcintyre RS, et al. Gut microbiota, bacterial translocation, and interactions with diet: Pathophysiological links between major depressive disorder and non-communicable medical comorbidities. Psychotherapy and Psychosomatics. 2017;86:31-46
  125. 125. Guillemot-Legris O, Muccioli GG. Obesity-induced neuroinflammation: Beyond the hypothalamus. Trends in Neurosciences. 2017;40:237-253
  126. 126. Novo AM, Batista S. Multiple sclerosis: Implications of obesity in neuroinflammation. Obesity and Brain Function. 2017:191-210
  127. 127. Segarra M, Aburto MR, Acker-Palmer A. Blood–brain barrier dynamics to maintain brain homeostasis. Trends in Neurosciences. 2021;44:393-405
  128. 128. Parimisetty A, Dorsemans A-C, Awada R, Ravanan P, Diotel N, LEfebvre D’Hellencourt, C. Secret talk between adipose tissue and central nervous system via secreted factors—An emerging frontier in the neurodegenerative research. Journal of Neuroinflammation. 2016;13:1-13
  129. 129. Stolp H, Dziegielewska K. Role of developmental inflammation and blood–brain barrier dysfunction in neurodevelopmental and neurodegenerative diseases. Neuropathology and Applied Neurobiology. 2009;35:132-146
  130. 130. Pugazhenthi S, Qin L, Reddy PH. Common neurodegenerative pathways in obesity, diabetes, and Alzheimer’s disease. Biochimica et Biophysica Acta (BBA)-Molecular Basis of Disease. 2017;1863:1037-1045
  131. 131. Spielman LJ, Little JP, Klegeris A. Inflammation and insulin/IGF-1 resistance as the possible link between obesity and neurodegeneration. Journal of Neuroimmunology. 2014;273:8-21
  132. 132. Miller AA, Spencer SJ. Obesity and neuroinflammation: A pathway to cognitive impairment. Brain, Behavior, and Immunity. 2014;42:10-21
  133. 133. Herradon G, Ramos-Alvarez MP, Gramage E. Connecting metainflammation and neuroinflammation through the PTN-MK-RPTPβ/ζ axis: Relevance in therapeutic development. Frontiers in Pharmacology. 2019;10:377
  134. 134. Ajami B, Bennett JL, Krieger C, Tetzlaff W, Rossi F. Local self-renewal can sustain CNS microglia maintenance and function throughout adult life. Nature Neuroscience. 2007;10:1538-1543
  135. 135. Keshk WA, Ibrahim MA, Shalaby SM, Zalat ZA, Elseady WS. Redox status, inflammation, necroptosis and inflammasome as indispensable contributors to high fat diet (HFD)-induced neurodegeneration; effect of N-acetylcysteine (NAC). Archives of Biochemistry and Biophysics. 2020;680:108227
  136. 136. Daulatzai MA. Cerebral hypoperfusion and glucose hypometabolism: Key pathophysiological modulators promote neurodegeneration, cognitive impairment, and Alzheimer’s disease. Journal of Neuroscience Research. 2017;95:943-972
  137. 137. Dey A, Allen JN, Fraser JW, Snyder LM, Tian Y, Zhang L, et al. Neuroprotective role of the Ron receptor tyrosine kinase underlying central nervous system inflammation in health and disease. Frontiers in Immunology. 2018;9:513
  138. 138. Fricker M, Tolkovsky AM, Borutaite V, Coleman M, Brown GC. Neuronal cell death. Physiological Reviews. 2018;98:813-880
  139. 139. Jan R. Understanding apoptosis and apoptotic pathways targeted cancer therapeutics. Advanced Pharmaceutical Bulletin. 2019;9:205
  140. 140. Xu X, Lai Y, Hua Z-C. Apoptosis and apoptotic body: Disease message and therapeutic target potentials. Bioscience Reports. 2019;39
  141. 141. Green DR, Llambi F. Cell death signaling. Cold Spring Harbor Perspectives in Biology. 2015;7:a006080
  142. 142. Kumari A, Kumar P, Yadav A. Impact of physical activity on central nervous system, neurogenesis and brains aging. Journal for Reattach Therapy and Developmental Diversities. 2022;5:229-234
  143. 143. Altman J. Are new neurons formed in the brains of adult mammals? Science. 1962;135:1127-1128
  144. 144. Altman J, Das GD. Post-natal origin of microneurones in the rat brain. Nature. 1965;207:953-956
  145. 145. Rakic P. Limits of neurogenesis in primates. Science. 1985;227:1054-1056
  146. 146. Gould E, Reeves AJ, Fallah M, Tanapat P, Gross CG, Fuchs E. Hippocampal neurogenesis in adult Old World primates. Proceedings of the National Academy of Sciences. 1999b;96:5263-5267
  147. 147. Kornack DR, Rakic P. Continuation of neurogenesis in the hippocampus of the adult macaque monkey. Proceedings of the National Academy of Sciences. 1999;96:5768-5773
  148. 148. Cameron HA, Mckay R. Discussion point stem cells and neurogenesis in the adult brain. Current Opinion in Neurobiology. 1998;8:677-680
  149. 149. Gage FH. Mammalian neural stem cells. Science. 2000;287:1433-1438
  150. 150. Reynolds BA, Weiss S. Generation of neurons and astrocytes from isolated cells of the adult mammalian central nervous system. Science. 1992;255:1707-1710
  151. 151. Alvarez-Buylla A, Lim DA. For the long run: Maintaining germinal niches in the adult brain. Neuron. 2004;41:683-686
  152. 152. Emsley JG, Mitchell BD, Kempermann G, Macklis JD. Adult neurogenesis and repair of the adult CNS with neural progenitors, precursors, and stem cells. Progress in Neurobiology. 2005;75:321-341
  153. 153. Li J, Tang Y, Cai D. IKKβ/NF-κB disrupts adult hypothalamic neural stem cells to mediate a neurodegenerative mechanism of dietary obesity and pre-diabetes. Nature Cell Biology. 2012;14:999-1012
  154. 154. Song H-J, Stevens CF, Gage FH. Neural stem cells from adult hippocampus develop essential properties of functional CNS neurons. Nature Neuroscience. 2002;5:438-445
  155. 155. Cai D. One Step from Prediabetes to Diabetes: Hypothalamic Inflammation? Oxford University Press; 2012
  156. 156. Cai D, Liu T. Hypothalamic inflammation: A double-edged sword to nutritional diseases. Annals of the New York Academy of Sciences. 2011;1243:E1-E39
  157. 157. Purkayastha S, Cai D. Disruption of neurogenesis by hypothalamic inflammation in obesity or aging. Reviews in Endocrine and Metabolic Disorders. 2013;14:351-356
  158. 158. Zhang G, Li J, Purkayastha S, Tang Y, Zhang H, Yin Y, et al. Hypothalamic programming of systemic ageing involving IKK-β, NF-κB and GnRH. Nature. 2013;497:211-216
  159. 159. Ekdahl CT, Claasen J-H, Bonde S, Kokaia Z, Lindvall O. Inflammation is detrimental for neurogenesis in adult brain. Proceedings of the National Academy of Sciences. 2003;100:13632-13637
  160. 160. Eikelenboom P, Veerhuis R, Scheper W, Rozemuller A, Van Gool WA, Hoozemans J. The significance of neuroinflammation in understanding Alzheimer’s disease. Journal of Neural Transmission. 2006;113:1685-1695
  161. 161. Hensley K, Mhatre M, Mou S, Pye QN, Stewart C, West M, et al. On the relation of oxidative stress to neuroinflammation: Lessons learned from the G93A-SOD1 mouse model of amyotrophic lateral sclerosis. Antioxidants & Redox Signaling. 2006;8:2075-2087
  162. 162. Minghetti L. Role of inflammation in neurodegenerative diseases. Current Opinion in Neurology. 2005;18:315-321
  163. 163. Taupin P. Adult neurogenesis, neuroinflammation and therapeutic potential of adult neural stem cells. International Journal of Medical Sciences. 2008;5:127
  164. 164. Posey KA, Clegg DJ, Printz RL, Byun J, Morton GJ, Vivekanandan-Giri A, et al. Hypothalamic proinflammatory lipid accumulation, inflammation, and insulin resistance in rats fed a high-fat diet. American Journal of Physiology-Endocrinology and Metabolism. 2009;296:E1003-E1012
  165. 165. Kleinridders A, Schenten D, Könner AC, Belgardt BF, Mauer J, Okamura T, et al. MyD88 signaling in the CNS is required for development of fatty acid-induced leptin resistance and diet-induced obesity. Cell Metabolism. 2009;10:249-259
  166. 166. Meng Q , Cai D. Defective hypothalamic autophagy directs the central pathogenesis of obesity via the IκB kinase β (IKKβ)/NF-κB pathway. Journal of Biological Chemistry. 2011;286:32324-32332
  167. 167. Milanski M, Degasperi G, Coope A, Morari J, Denis R, Cintra DE, et al. Saturated fatty acids produce an inflammatory response predominantly through the activation of TLR4 signaling in hypothalamus: Implications for the pathogenesis of obesity. Journal of Neuroscience. 2009;29:359-370
  168. 168. Oh IS, Thaler JP, Ogimoto K, Wisse BE, Morton GJ, Schwartz MW. Central administration of interleukin-4 exacerbates hypothalamic inflammation and weight gain during high-fat feeding. American Journal of Physiology-Endocrinology and Metabolism. 2010;299:E47-E53
  169. 169. Purkayastha S, Zhang H, Zhang G, Ahmed Z, Wang Y, Cai D. Neural dysregulation of peripheral insulin action and blood pressure by brain endoplasmic reticulum stress. Proceedings of the National Academy of Sciences. 2011;108:2939-2944
  170. 170. Ballard C, Gauthier S, Corbett A, Brayne C, Aarsland D, Jones E. Alzheimer’s disease. The Lancet. 2011;377:1019-1031
  171. 171. Beydoun MA, Beydoun H, Wang Y. Obesity and central obesity as risk factors for incident dementia and its subtypes: A systematic review and meta-analysis. Obesity Reviews. 2008;9:204-218
  172. 172. Hamer M, Chida Y. Physical activity and risk of neurodegenerative disease: A systematic review of prospective evidence. Psychological Medicine. 2009;39:3-11
  173. 173. Lu F-P, Lin K-P, Kuo H-K. Diabetes and the risk of multi-system aging phenotypes: A systematic review and meta-analysis. PLoS One. 2009;4:e4144
  174. 174. Grote HE, Hannan AJ. Regulators of adult neurogenesis in the healthy and diseased brain. Clinical and Experimental Pharmacology and Physiology. 2007;34:533-545
  175. 175. Monje ML, Toda H, Palmer TD. Inflammatory blockade restores adult hippocampal neurogenesis. Science. 2003;302:1760-1765
  176. 176. Lee DA, Bedont JL, Pak T, Wang H, Song J, Miranda-Angulo A, et al. Tanycytes of the hypothalamic median eminence form a diet-responsive neurogenic niche. Nature Neuroscience. 2012;15:700-702
  177. 177. Mcnay DE, Briançon N, Kokoeva MV, Maratos-Flier E, Flier JS. Remodeling of the arcuate nucleus energy-balance circuit is inhibited in obese mice. The Journal of Clinical Investigation. 2012;122:142-152
  178. 178. Drapeau E, Montaron M-F, Aguerre S, Abrous DN. Learning-induced survival of new neurons depends on the cognitive status of aged rats. Journal of Neuroscience. 2007;27:6037-6044
  179. 179. Gould E, Beylin A, Tanapat P, Reeves A, Shors TJ. Learning enhances adult neurogenesis in the hippocampal formation. Nature Neuroscience. 1999a;2:260-265
  180. 180. Shors TJ, Miesegaes G, Beylin A, Zhao M, Rydel T, Gould E. Neurogenesis in the adult is involved in the formation of trace memories. Nature. 2001;410:372-376
  181. 181. Leal G, Bramham C, Duarte C. BDNF and hippocampal synaptic plasticity. Vitamins and Hormones. 2017;104:153-195
  182. 182. Reichelt AC, Hare DJ, Bussey TJ, Saksida LM. Perineuronal nets: Plasticity, protection, and therapeutic potential. Trends in Neurosciences. 2019;42:458-470
  183. 183. Cui Y, Cao K, Lin H, Cui S, Shen C, Wen W, et al. Early-life stress induces depression-like behavior and synaptic-plasticity changes in a maternal separation rat model: Gender difference and metabolomics study. Frontiers in Pharmacology. 2020;11:102
  184. 184. Jackson J, Jambrina E, Li J, Marston H, Menzies F, Phillips K, et al. Targeting the synapse in Alzheimer’s disease. Frontiers in Neuroscience. 2019;13:735
  185. 185. Ramiro-Cortés Y, Hobbiss AF, Israely I. Synaptic competition in structural plasticity and cognitive function. Philosophical Transactions of the Royal Society B: Biological Sciences. 2014;369:20130157
  186. 186. Abidin İ, Aydin-Abidin S, Bodur A, Ince İ, Alver A. Brain-derived neurotropic factor (BDNF) heterozygous mice are more susceptible to synaptic protein loss in cerebral cortex during high fat diet. Archives of Physiology and Biochemistry. 2018;124:442-447
  187. 187. Arnold SE, Lucki I, Brookshire BR, Carlson GC, Browne CA, Kazi H, et al. High fat diet produces brain insulin resistance, synaptodendritic abnormalities and altered behavior in mice. Neurobiology of Disease. 2014;67:79-87
  188. 188. Bhat NR, Thirumangalakudi L. Increased tau phosphorylation and impaired brain insulin/IGF signaling in mice fed a high fat/high cholesterol diet. Journal of Alzheimer’s Disease. 2013;36:781-789
  189. 189. Bocarsly ME, Fasolino M, Kane GA, Lamarca EA, Kirschen GW, Karatsoreos IN, et al. Obesity diminishes synaptic markers, alters microglial morphology, and impairs cognitive function. Proceedings of the National Academy of Sciences. 2015;112:15731-15736
  190. 190. Cavaliere G, Viggiano E, Trinchese G, De Filippo C, Messina A, Monda V, et al. Long feeding high-fat diet induces hypothalamic oxidative stress and inflammation, and prolonged hypothalamic AMPK activation in rat animal model. Frontiers in Physiology. 2018;9:818
  191. 191. Hao S, Dey A, Yu X, Stranahan AM. Dietary obesity reversibly induces synaptic stripping by microglia and impairs hippocampal plasticity. Brain, Behavior, and Immunity. 2016;51:230-239
  192. 192. Papazoglou IK, Jean A, Gertler A, Taouis M, Vacher C-M. Hippocampal GSK3β as a molecular link between obesity and depression. Molecular Neurobiology. 2015;52:363-374
  193. 193. Puig KL, Floden AM, Adhikari R, Golovko MY, Combs CK. Amyloid precursor protein and proinflammatory changes are regulated in brain and adipose tissue in a murine model of high fat diet-induced obesity. PLoS One. 2012;7:e30378
  194. 194. Valcarcel-Ares MN, Tucsek Z, Kiss T, Giles CB, Tarantini S, Yabluchanskiy A, et al. Obesity in aging exacerbates neuroinflammation, dysregulating synaptic function-related genes and altering eicosanoid synthesis in the mouse hippocampus: Potential role in impaired synaptic plasticity and cognitive decline. The Journals of Gerontology: Series A. 2019;74:290-298
  195. 195. Dutheil S, Ota KT, Wohleb ES, Rasmussen K, Duman RS. High-fat diet induced anxiety and anhedonia: Impact on brain homeostasis and inflammation. Neuropsychopharmacology. 2016;41:1874-1887
  196. 196. Salameh TS, Mortell WG, Logsdon AF, Butterfield DA, Banks WA. Disruption of the hippocampal and hypothalamic blood–brain barrier in a diet-induced obese model of type II diabetes: Prevention and treatment by the mitochondrial carbonic anhydrase inhibitor, topiramate. Fluids and Barriers of the CNS. 2019;16:1-17
  197. 197. Tongiorgi E. Activity-dependent expression of brain-derived neurotrophic factor in dendrites: Facts and open questions. Neuroscience Research. 2008;61:335-346
  198. 198. Sharma S, Fulton S. Diet-induced obesity promotes depressive-like behaviour that is associated with neural adaptations in brain reward circuitry. International Journal of Obesity. 2013;37:382-389
  199. 199. Notaras M, van den Buuse M. Brain-derived neurotrophic factor (BDNF): Novel insights into regulation and genetic variation. The Neuroscientist. 2019;25:434-454
  200. 200. Vanevski F, Xu B. Molecular and neural bases underlying roles of BDNF in the control of body weight. Frontiers in Neuroscience. 2013;7:37
  201. 201. Gupta S, Mishra R, Kusum S, Spedding M, Meiri K, Gressens P, et al. GAP-43 is essential for the neurotrophic effects of BDNF and positive AMPA receptor modulator S18986. Cell Death & Differentiation. 2009;16:624-637
  202. 202. Wang Z, Ge Q , Wu Y, Zhang J, Gu Q , Han J. Impairment of long-term memory by a short-term high-fat diet via hippocampal oxidative stress and alterations in synaptic plasticity. Neuroscience. 2020;424:24-33
  203. 203. 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. European Journal of Neuroscience. 2004;19:1699-1707
  204. 204. Jovanovic J, Czernik AJ, Fienberg AA, Greengard P, Sihra TS. Synapsins as mediators of BDNF-enhanced neurotransmitter release. Nature Neuroscience. 2000;3:323-329
  205. 205. Pandit M, Behl T, Sachdeva M, Arora S. Role of brain derived neurotropic factor in obesity. Obesity Medicine. 2020;17:100189
  206. 206. Criscuolo C, Fabiani C, Bonadonna C, Origlia N, Domenici L. BDNF prevents amyloid-dependent impairment of LTP in the entorhinal cortex by attenuating p38 MAPK phosphorylation. Neurobiology of Aging. 2015;36:1303-1309
  207. 207. Kurhe Y, Mahesh R, Devadoss T. Novel 5-HT3 receptor antagonist QCM-4 attenuates depressive-like phenotype associated with obesity in high-fat-diet-fed mice. Psychopharmacology. 2017;234:1165-1179
  208. 208. Park HR, Park M, Choi J, Park K-Y, Chung HY, Lee J. A high-fat diet impairs neurogenesis: Involvement of lipid peroxidation and brain-derived neurotrophic factor. Neuroscience Letters. 2010;482:235-239
  209. 209. Reichelt AC, Maniam J, Westbrook RF, Morris MJ. Dietary-induced obesity disrupts trace fear conditioning and decreases hippocampal reelin expression. Brain, Behavior, and Immunity. 2015;43:68-75
  210. 210. Jin Y, Sun LH, Yang W, Cui RJ, Xu SB. The role of BDNF in the neuroimmune axis regulation of mood disorders. Frontiers in Neurology. 2019;10:515
  211. 211. Puig KL, Brose SA, Zhou X, Sens MA, Combs GF, Jensen MD, et al. Amyloid precursor protein modulates macrophage phenotype and diet-dependent weight gain. Scientific Reports. 2017;7:1-14
  212. 212. Mehta P, Pirttila T, Patrick B, Barshatzky M, Mehta S. Amyloid β protein 1-40 and 1-42 levels in matched cerebrospinal fluid and plasma from patients with Alzheimer disease. Neuroscience Letters. 2001;304:102-106
  213. 213. Ismail N, Ismail M, Azmi NH, Bakar MFA, Yida Z, Abdullah MA, et al. Thymoquinone-rich fraction nanoemulsion (TQRFNE) decreases Aβ40 and Aβ42 levels by modulating APP processing, up-regulating IDE and LRP1, and down-regulating BACE1 and RAGE in response to high fat/cholesterol diet-induced rats. Biomedicine & Pharmacotherapy. 2017;95:780-788
  214. 214. Nuzzo D, Baldassano S, Amato A, Picone P, Galizzi G, Caldara GF, et al. Glucagon-like peptide-2 reduces the obesity-associated inflammation in the brain. Neurobiology of Disease. 2019;121:296-304
  215. 215. Lee YH, Tharp WG, Maple RL, Nair S, Permana PA, Pratley RE. Amyloid precursor protein expression is upregulated in adipocytes in obesity. Obesity. 2008;16:1493-1500
  216. 216. Sommer G, Kralisch S, Lipfert J, Weise S, Krause K, Jessnitzer B, et al. Amyloid precursor protein expression is induced by tumor necrosis factor α in 3T3-L1 adipocytes. Journal of Cellular Biochemistry. 2009;108:1418-1422
  217. 217. Tarantini S, Valcarcel-Ares MN, Yabluchanskiy A, Tucsek Z, Hertelendy P, Kiss T, et al. Nrf2 deficiency exacerbates obesity-induced oxidative stress, neurovascular dysfunction, blood–brain barrier disruption, neuroinflammation, amyloidogenic gene expression, and cognitive decline in mice, mimicking the aging phenotype. The Journals of Gerontology: Series A. 2018;73:853-863
  218. 218. Kohjima M, Sun Y, Chan L. Increased food intake leads to obesity and insulin resistance in the tg2576 Alzheimer’s disease mouse model. Endocrinology. 2010;151:1532-1540
  219. 219. Thirumangalakudi L, Prakasam A, Zhang R, Bimonte-Nelson H, Sambamurti K, Kindy MS, et al. High cholesterol-induced neuroinflammation and amyloid precursor protein processing correlate with loss of working memory in mice. Journal of Neurochemistry. 2008;106:475-485
  220. 220. Nakandakari SCBR, Munoz VR, Kuga GK, Gaspar RC, Sant’ana MR, Pavan ICB, et al. Short-term high-fat diet modulates several inflammatory, ER stress, and apoptosis markers in the hippocampus of young mice. Brain, Behavior, and Immunity. 2019;79:284-293
  221. 221. Frühbeis C, Fröhlich D, Kuo WP, Krämer-Albers E-M. Extracellular vesicles as mediators of neuron-glia communication. Frontiers in Cellular Neuroscience. 2013;7:182
  222. 222. Duffy C, Hofmeister J, Nixon JP, Butterick T. High fat diet increases cognitive decline and neuroinflammation in a model of orexin loss. Neurobiology of Learning and Memory. 2019;157:41-47
  223. 223. Vinuesa A, Bentivegna M, Calfa G, Filipello F, Pomilio C, Bonaventura MM, et al. Early exposure to a high-fat diet impacts on hippocampal plasticity: Implication of microglia-derived exosome-like extracellular vesicles. Molecular Neurobiology. 2019;56:5075-5094
  224. 224. Cope EC, Lamarca EA, Monari PK, Olson LB, Martinez S, Zych AD, et al. Microglia play an active role in obesity-associated cognitive decline. Journal of Neuroscience. 2018;38:8889-8904
  225. 225. Battú CE, Rieger D, Loureiro S, Furtado GV, Bock H, Saraiva-Pereira M-L, et al. Alterations of PI3K and Akt signaling pathways in the hippocampus and hypothalamus of Wistar rats treated with highly palatable food. Nutritional Neuroscience. 2012;15:10-17
  226. 226. Kothari V, Luo Y, Tornabene T, O’Neill AM, Greene MW, Geetha T, et al. High fat diet induces brain insulin resistance and cognitive impairment in mice. Biochimica et Biophysica Acta (BBA)-Molecular Basis of Disease. 2017;1863:499-508
  227. 227. Wijtenburg SA, Kapogiannis D, Korenic SA, Mullins RJ, Tran J, Gaston FE, et al. Brain insulin resistance and altered brain glucose are related to memory impairments in schizophrenia. Schizophrenia Research. 2019;208:324-330
  228. 228. Matsuzaki T, Sasaki K, Tanizaki Y, Hata J, Fujimi K, Matsui Y, et al. Insulin resistance is associated with the pathology of Alzheimer disease: The Hisayama study. Neurology. 2010;75:764-770
  229. 229. Lomniczi A, Ojeda SR. A role for glial cells of the neuroendocrine brain in the central control of female sexual development. In: Astrocytes in (Patho) Physiology of the Nervous System. Springer; 2009
  230. 230. Boeckers TM, Bockmann J, Kreutz MR, Gundelfinger ED. ProSAP/shank proteins–a family of higher order organizing molecules of the postsynaptic density with an emerging role in human neurological disease. Journal of Neurochemistry. 2002;81:903-910
  231. 231. Mielke JG, Nicolitch K, Avellaneda V, Earlam K, Ahuja T, Mealing G, et al. Longitudinal study of the effects of a high-fat diet on glucose regulation, hippocampal function, and cerebral insulin sensitivity in C57BL/6 mice. Behavioural Brain Research. 2006;175:374-382
  232. 232. Vander Heide LP, Kamal A, Artola A, Gispen WH, Ramakers GM. Insulin modulates hippocampal activity-dependent synaptic plasticity in a N-methyl-d-aspartate receptor and phosphatidyl-inositol-3-kinase-dependent manner. Journal of Neurochemistry. 2005;94:1158-1166
  233. 233. Anfal A-D, Ahram M, Hayder A. Effects of obesity on hippocampus function: Synaptic plasticity hypothesis. Obesity Medicine. 2020;19:100246
  234. 234. Dingess PM, Darling RA, Kurt Dolence E, Culver BW, Brown TE. Exposure to a diet high in fat attenuates dendritic spine density in the medial prefrontal cortex. Brain Structure and Function. 2017;222:1077-1085
  235. 235. Uranga RM, Bruce-Keller AJ, Morrison CD, Fernandez-Kim SO, Ebenezer PJ, Zhang L, et al. Intersection between metabolic dysfunction, high fat diet consumption, and brain aging. Journal of Neurochemistry. 2010;114:344-361
  236. 236. Magariños AM, Mcewen BS. Experimental diabetes in rats causes hippocampal dendritic and synaptic reorganization and increased glucocorticoid reactivity to stress. Proceedings of the National Academy of Sciences. 2000;97:11056-11061
  237. 237. Pamidi N, Satheesha Nayak BN. Effect of streptozotocin induced diabetes on rat hippocampus. Bratislavské Lekárske Listy. 2012;113:583-588
  238. 238. Ramos-Rodriguez JJ, Molina-Gil S, Ortiz-Barajas O, Jimenez-Palomares M, Perdomo G, Cozar-Castellano I, et al. Central proliferation and neurogenesis is impaired in type 2 diabetes and prediabetes animal models. PLoS One. 2014;9:e89229
  239. 239. Stranahan AM, Norman ED, Lee K, Cutler RG, Telljohann RS, Egan JM, et al. Diet-induced insulin resistance impairs hippocampal synaptic plasticity and cognition in middle-aged rats. Hippocampus. 2008;18:1085-1088
  240. 240. Karimi SA, Salehi I, Komaki A, Sarihi A, Zarei M, Shahidi S. Effect of high-fat diet and antioxidants on hippocampal long-term potentiation in rats: An in vivo study. Brain Research. 2013;1539:1-6
  241. 241. Yoon G, Cho KA, Song J, Kim Y-K. Transcriptomic analysis of high fat diet fed mouse brain cortex. Frontiers in Genetics. 2019;10:83
  242. 242. Lindqvist A, Mohapel P, Bouter B, Frielingsdorf H, Pizzo D, Brundin P, et al. High-fat diet impairs hippocampal neurogenesis in male rats. European Journal of Neurology. 2006;13:1385-1388

Written By

Amina Khatun, Surendra Patra, Kuntal Ghosh, Shrabani Pradhan and Sudipta Chakrabarti

Submitted: 22 November 2022 Reviewed: 20 February 2023 Published: 31 March 2023