Next Article in Journal
Overweight or Obesity Rate and Risk Factors in First-Episode and Drug-Naïve Patients with Major Depressive Disorder with Comorbid Abnormal Lipid Metabolism: A Large-Scale Cross-Sectional Study
Previous Article in Journal
Bayesian 13C-Metabolic Flux Analysis of Parallel Tracer Experiments in Granulocytes: A Directional Shift within the Non-Oxidative Pentose Phosphate Pathway Supports Phagocytosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Potential Role for the Ketogenic Diet in Alzheimer’s Disease Treatment: Exploring Pre-Clinical and Clinical Evidence

by
Tadeu P. D. Oliveira
1,
Ana L. B. Morais
2,
Pedro L. B. dos Reis
2,
András Palotás
3,4,5 and
Luciene B. Vieira
2,*
1
Departamento de Fisiologia e Centro de Investigação em Medicina Molecular (CIMUS), Universidad De Santiago de Compostela, 15782 Santiago de Compostela, Spain
2
Departamento de Farmacologia, Instituto de Ciências Biológicas (ICB), Universidade Federal de Minas Gerais, Belo Horizonte 31270-901, Brazil
3
Asklepios-Med (Private Medical Practice and Research Center), H-6722 Szeged, Hungary
4
Kazan Federal University, Kazan R-420012, Russia
5
Tokaj-Hegyalja University, H-3910 Tokaj, Hungary
*
Author to whom correspondence should be addressed.
Metabolites 2024, 14(1), 25; https://doi.org/10.3390/metabo14010025
Submission received: 22 November 2023 / Revised: 18 December 2023 / Accepted: 21 December 2023 / Published: 29 December 2023
(This article belongs to the Special Issue Brain Metabolic Alterations in Neurodegenerative Diseases)

Abstract

:
Given the remarkable progress in global health and overall quality of life, the significant rise in life expectancy has become intertwined with the surging occurrence of neurodegenerative disorders (NDs). This emerging trend is poised to pose a substantial challenge to the fields of medicine and public health in the years ahead. In this context, Alzheimer’s disease (AD) is regarded as an ND that causes recent memory loss, motor impairment and cognitive deficits. AD is the most common cause of dementia in the elderly and its development is linked to multifactorial interactions between the environment, genetics, aging and lifestyle. The pathological hallmarks in AD are the accumulation of β-amyloid peptide (Aβ), the hyperphosphorylation of tau protein, neurotoxic events and impaired glucose metabolism. Due to pharmacological limitations and in view of the prevailing glycemic hypometabolism, the ketogenic diet (KD) emerges as a promising non-pharmacological possibility for managing AD, an approach that has already demonstrated efficacy in addressing other disorders, notably epilepsy. The KD consists of a food regimen in which carbohydrate intake is discouraged at the expense of increased lipid consumption, inducing metabolic ketosis whereby the main source of energy becomes ketone bodies instead of glucose. Thus, under these dietary conditions, neuronal death via lack of energy would be decreased, inasmuch as the metabolism of lipids is not impaired in AD. In this way, the clinical picture of patients with AD would potentially improve via the slowing down of symptoms and delaying of the progression of the disease. Hence, this review aims to explore the rationale behind utilizing the KD in AD treatment while emphasizing the metabolic interplay between the KD and the improvement of AD indicators, drawing insights from both preclinical and clinical investigations. Via a comprehensive examination of the studies detailed in this review, it is evident that the KD emerges as a promising alternative for managing AD. Moreover, its efficacy is notably enhanced when dietary composition is modified, thereby opening up innovative avenues for decreasing the progression of AD.

Graphical Abstract

1. Review Design

The current narrative review was collaboratively designed by five contributors to plan the research question, drawing insights from considerations outlined in the abstract: “Ketogenic dietary therapies for neurodegenerative disease and Alzheimer’s disease”. Our research strategy was devised for PubMed and Scopus, aligning with predefined keywords. These keywords were employed both independently and in combination, utilizing the Boolean AND operator to establish logical relationships among concepts. The research approach involved advanced search techniques, and specific limits were set: papers published until July 2023 encompassing studies involving humans and rodent models, with a focus on English-language publications.

2. Epidemiology, Neuropathological Insights and Symptoms in Alzheimer ’s Disease

Alzheimer’s disease (AD) is the most prevalent form of dementia [1,2,3,4]. It is also considered to be a growing burden for the health care system and is one of the most expensive chronic diseases of old age [3,5,6,7]. In 2010, a total of 604 billion USD was spent on AD globally; these direct medical costs are exacerbated by lost productivity attributed to this debilitating disease. Projections for the year 2050 estimate that such costs would surpass 1 trillion USD in total [6,8,9,10]. These economic expenses are paralleled with the incidence of patients with Alzheimer’s, several of which are forecasted to rise above 150 million by 2050 [10,11]. In addition, AD has an intense impact on the individual and social spheres, with caregivers of affected people being more vulnerable to anxiety and depression [5,12]. Therefore, the functional capacity and autonomy of the elderly may be more important than mortality, as they relate to quality of life [6]. In all populations, AD is prevalent at advanced ages, notably between 70 and 80 years [2,13]. However, differences were found in the association of AD among ethnic groups and sex [14,15]. Importantly, a greater susceptibility to developing AD has been reported in Caucasians, Japanese, and women, as nearly two-thirds of people diagnosed are women [16,17].
Neuropathologically, AD is characterized by the deposition of β-amyloid peptides (Aβ) and the accumulation of neurofibrillary tangles (NFTs) composed of hyper-phosphorylated tau protein. Under physiological conditions, tau protein exists as a soluble and unfolded protein that interacts with tubulin to promote the assembly and stabilization of microtubules [18,19]. Aβ peptides can self-aggregate into soluble oligomers, or into insoluble fibers, forming the amyloid plaques. The Aβ from 36 to 43 amino acids is a product of proteolytic processing of protein amyloid precursor (APP), a type I transmembrane protein [20]. The cleavage of APP by the enzyme β-secretase (BACE-1) generates a C-terminal fragment (β-CTF), which is an immediate substrate for the γ-secretase enzyme that cleaves β-CTF to produce a spectrum of Aβ peptides with different lengths, i.e., the amyloidogenic pathway [21]. Conversely, when APP is cleaved by the α-secretase enzyme, it generates a C-terminus that subsequently undergoes cleavage by γ-secretase. This pathway, known as the non-amyloidogenic pathway, leads to the production of a shorter APP fragment (p3). This seemingly innocuous fragment may, in fact, hold pivotal roles in various processes, including cell growth, adhesion, synaptic plasticity, and the regulation of metal ion homeostasis [22,23].
Moreover, the amyloid hypothesis postulates that the accumulation of Aβ leads to a neurodegenerative cascade, resulting in synaptic dysfunction, NFT formation and, ultimately, neuronal loss in susceptible brain regions. Thus, AD is associated with neuronal degeneration, manifesting selective impairments in the hippocampus and neocortex regions, predominantly affecting cholinergic neuronal populations, and resulting in neurological impairments [24,25,26]. While the manifestation of symptoms may exhibit heterogeneity, AD is characterized by a progressive and irreversible decline in memory, cognition, thinking and language [27,28,29]. This decline culminates in diminished independence in performing daily tasks, accompanied by behavioral and motor impairments, as well as escalating psychiatric symptoms. These challenges tend to intensify as the neurodegenerative syndrome progresses [16,24,30]. Moreover, shared mechanisms of cell death in AD are related to increased oxidative stress, mitochondrial injury, hypometabolism, disruption of the blood–brain barrier (BBB) and neuroinflammation [31,32,33]. In the latter process, the increased formation of Aβ and NFTs is associated with the presence of activated microglia, cytokines, cyclooxygenase 2 (COX-2) and other inflammatory substances [28,32].

3. Risk Factors for AD

Despite 115 years having passed since Alois Alzheimer first described the disease, our understanding of AD remains incomplete. Consequently, researchers have explored genetic and environmental factors as potentially influential elements in the development of this condition [34,35]. Genetic risk factors for AD include mutations in amyloid precursor protein (APP), presenilin 1 (PSEN1), presenilin 2 (PSEN2) and apolipoprotein E4 (APOE4). Nonetheless, APOE4 stands as the most extensively established genetic risk factor for AD susceptibility beyond the age of 65 years [36,37,38,39]. It is linked to an elevated risk of NFT formation and a reduction in Aβ clearance, resulting in the accumulation of neurotoxic fragments [37,38]. Furthermore, while genes play a role in the risk of developing AD, they may account for only a modest portion of that risk [2,36]. It is important to highlight that the risk factors identified for AD share features and are not independent [40]. Notably, some of these factors encompass treatable medical conditions like stroke, hypertension and diabetes [36]. Epidemiologic studies revealed that the risk of AD is increased by 50–100% by type 2 diabetes mellitus (T2DM). A notable observation is that the APOE4 gene serves as a risk factor for the development of type 2 diabetes mellitus (T2DM), although it is important to recognize that not all individuals with T2DM possess the APOE4 gene. Additionally, this apolipoprotein has been linked to a reduction in glucose consumption by the brain [41]. Thus, T2DM is linked to an increased risk of developing late-onset AD (LOAD) [42,43]. Regarding modifiable environmental risks, numerous reviews have emphasized the substantial evidence associating smoking [42], nutrition [44] and obesity [45], and such findings demonstrate that behavioral aspects can influence the onset of clinical manifestations of AD [46,47,48].

4. Pharmacological Treatment for AD

Currently, there is neither a cure for AD [49] nor an effective drug for prevention or treatment that modifies the progression of the disease [41,50]. Given the severity of AD, advances in pharmacological treatments are progressing slowly [51], as only a few drugs are approved. The drug therapy for AD is symptomatic or palliative and not effective for advanced stages of the disease [49,52,53,54]. Such pharmacological treatments are focused on improving cholinergic transmission, and they are divided by the mechanism of action into two classes: cholinesterase inhibitors (ChEIs), used for mild to moderate stages, and an NMDA receptor antagonist (N-Methyl-D-aspartate), Memantine [49], used for moderate to severe stages or in cases of intolerance and contraindication [8,55,56]. ChEIs have moderate symptomatic benefits regarding cognition, functionality and behavior [36,57,58]. This class includes Donepezil, Galantamine and Rivastigmine [59]. Nevertheless, these drugs exhibit variations in certain pharmacological characteristics: Donepezil and Rivastigmine boast longer half-lives, but, in addition to acetylcholinesterase inhibition, Rivastigmine also deactivates butyrylcholinesterase. On the other hand, Galantamine demonstrates an additional enhancement in nicotinic receptor transmission [60,61]. As for the NMDA receptor antagonist, Memantine improves cognition, functionality and the management of agitation and aggression [36,58,62]. Furthermore, combining Memantine with Donepezil can lead to improved patient outcomes [61].
Additional treatment options are non-pharmacological, often more cost-effective and dependent on human effort [63]. This category encompasses numerous suggested approaches, including socialization [64], cognitive training [65], calorie restriction and exercise [8,60]. Among these alternatives, the ketogenic diet (KD) is currently under investigation as an adjuvant therapy [66,67,68,69].

5. Ketogenic Diet and Ketone Body Biosynthesis

The KD is a diet based on reducing carbohydrate consumption and increasing lipid intake [70,71]. This leads to a decrease in the use of glucose, which is no longer the main energy source, promoting the use of ketone bodies (ketones) from the breakdown of fatty acids (FAs) [72,73,74]. The hepatic metabolism of FAs produces ketones commonly used as substrates for energy: acetoacetate, acetone and beta-hydroxybutyrate (BOHb) [75,76]. Normally, ketones are produced in starvation [77], fasting [78,79], prolonged physical exercises [80], pregnancy [81] and in diets with high-fat and low-carbohydrate rates [82]. Thus, when glucose stores in the body are low, more FAs are made available to the liver for oxidation, leading to the consequent production of energy-rich molecules, mainly acetyl-CoA. Acetyl-CoA can enter the citric acid cycle in the liver or be used for the synthesis of ketones. Once in the liver cells, the fatty acid will be directed to the mitochondrial matrix by carnitine palmitoyltransferase (CPT1/2) where it initially undergoes a β-oxidation generating Acetyl-CoA that will undergo the process of ketogenesis [83]. In sequence in the process, the thiolase-2 enzyme acts in the conversion of two molecules of Acetyl-CoA to Acetoacetyl-CoA (AcAc-CoA) [83,84]. This molecule undergoes catalysis by the enzyme 3-Hydroxymethyl glutaryl-CoA synthase 2 (HMGCS2), resulting in the generation of hydroxymethylglutaryl-CoA (HMG-CoA) [84]. Subsequently, HMG-CoA is converted into acetoacetate and Acetyl-CoA via the catalytic action of HMG-CoA lyase [85]. Furthermore, acetoacetate can be reduced to D-β-hydroxybutyrate (D-βOHB) or decarboxylated to acetone [83]. After their formation, ketones are released from cells by monocarboxylate transporters (MCT1/2) and fall into the bloodstream to reach extrahepatic tissues for terminal oxidation [83]. Through the same MCT1/2 channels, ketone bodies enter the mitochondrial matrix of cells where they undergo the action of Succinyl-CoA: 3-ketoacid CoA transferase (SCOT) that transfers the CoA portion of succinyl-CoA to form Acetoacetyl-CoA [86]. The final part of the inverse process leads to the formation of Acetyl-CoA that is introduced into the TCA cycle for the formation of ATP that is used as an energy source in cases of glucose deprivation [87] (Figure 1).

6. Types of KD

There are different types of KDs, listed as: classic long-chain triglyceride KD (LCT), medium-chain triglyceride KD (MCT), modified Atkins diet (MAD) and low glycemic index diet (LOGI) (Table 1) [66,88,89,90]. The four diets have the same original formula, characterized by a high rate of fat and low amount of carbohydrate in their composition. However, they have occasional variations in the composition weight and ingredient restrictions [90]. LCT offers around 90% of energy in the form of fat and 10% of carbohydrates and proteins [90]. The most recommended ratio is 4:1 to 3:1 (fats: proteins and carbohydrates), but the use of each diet can be evaluated based on the patient’s profile and the most appropriate type of diet [90]. The diet ratio represents the balance between fat and protein plus carbohydrate grams. For instance, a “1800 kcal 4:1 ratio classic KD” contains four times the grams of fat compared to protein. This ratio can be customized to enhance seizure management or to make it more accommodating for improved tolerance. In contrast to the standard KD, the MCTKD is not influenced by food ratios; instead, it depends on the proportion of calories derived from MCT oil as a crucial source of ketones [91].
MCT has a distinct composition, primarily comprising about 60% octanoic acid, an eight-carbon fatty acid, and roughly 40% decanoic acid, a ten-carbon fatty acid [92]. Unlike the conventional KD, which relies more on medium-chain fats for dietary energy, the MCT-based diet permits a broader inclusion of carbohydrates. Due to the swift metabolism of these shorter fatty acids, this distinction results in a more efficient synthesis of ketones [92]. The process starts with dietary triglycerides in the form of MCT supplements, undergoing breakdown in the gastrointestinal tract by specialized lipases with a preference for hydrolyzing medium-chain esters rather than long-chain esters. Consequently, MCTs are converted into medium-chain fatty acids, characterized by their carbon atom content ranging from six to twelve [93]. This unique property enables direct absorption through the intestinal wall, leading to swift transportation to the liver. Once in the liver, the medium-chain fatty acids, including decanoic acid and octanoic acid, undergo rapid metabolism via a process known as β-oxidation [93]. The integration of an MCT-rich diet with an elevated carbohydrate intake sets this approach apart from the conventional KD, offering a balanced and efficient means of achieving ketosis while harnessing the advantages of medium-chain fats [92].
Due to the highly restrictive dietary regimen and concerning side effects associated with KDs, their implementation in pediatric patients is challenging. In this context, the modified Atkins diet (MAD) emerges as a more balanced and easily applicable alternative dietary therapy. The advantages of increased tolerance and sustainable treatment approach could potentially position MAD as the preferred choice. Unlike the classic KD, MAD shares similar food choices but eliminates the need for precise ingredient weighing. It also deviates from a strict ketogenic ratio and lacks restrictions on protein, fluid and calories. MAD is used to treat some metabolic paroxysmal movement disorders, such as those observed in glucose transporter type 1 deficiency syndrome (GLUT1DS) [88,89,90].
Another type of diet that combines the principles of the Mediterranean diet with the macronutrient composition of a KD is the Mediterranean Ketogenic Diet (MKD). Inspired by the heart-healthy Mediterranean diet, the MKD incorporates high consumption of fruits, vegetables, whole grains and healthy fats, notably olive oil. Simultaneously, it aligns with the principles of a KD by emphasizing low carbohydrate intake and promoting a state of ketosis. The MKD includes the use of olive oil as a primary source of fat, moderate consumption of fish and poultry, and limited intake of red meat and processed foods. The MKD encompasses several common variations, including the Very Low-Calorie Ketogenic Diet (VLCKD) which is characterized by a low carbohydrate content (<50 g/day), 1–1.5 g of protein/kg of ideal body weight, 15–30 g of fat/day and a daily intake of about 500–800 calories [94]. In addition, the High-Fat Ketogenic Diet (HFKD) is based on a higher proportion of daily calories sourced from fat, typically ranging between 75–80%. In contrast to the traditional KD, the HFKD allows for a modestly increased protein intake, contributing to a more flexible nutritional profile [94].

7. Possible Risks of KD

The benefits of ketones produced in the liver go beyond the energy supply of tissues such as the brain, skeletal muscle and heart [83]. Ketones antagonize inflammatory processes and oxidative stress [95,96], acting as signaling mediators [83]. Although the KD presents possible benefits to the organism [97], its use can also promote adverse effects such as headache, gastrointestinal pain, constipation, nausea, fatty diarrhea, fatigue, vomiting and other gastrointestinal problems [98,99,100]. These symptoms are usually associated with acute use of the KD as reported in studies with young and adult patients [99,100]. All the symptoms caused by the KD in the first few days are usually called “keto flu” [100]. It has been shown that the acute symptoms pass after a short period of time and patients who use the KD for more than one year can report different types of symptoms, as vitamin and mineral deficiencies, kidney stones, hyperuricemia, lethargy and infectious diseases, which can be harmful [99].
In a study involving obese patients, the impact of a KD was assessed over a 24-week period. The findings demonstrated that the KD treatment yielded positive outcomes, including significant weight loss and improvements in the patients’ lipid profiles, with no notable adverse effects reported [101]. However, it is crucial to emphasize that individuals with a genetic predisposition to cholesterol metabolism dysregulation may experience an exaggerated rise in cholesterol levels when adhering to a KD. Moreover, there remain certain uncertainties surrounding the prolonged use of this diet, primarily due to the limited evidence available for durations exceeding one year. These uncertainties encompass concerns related to cardiovascular risks and disruptions in lipid metabolism as well as impacts on hormone regulation, such as insulin [102,103]. Although there is evidence supporting KD effectiveness in specific contexts, such as obesity, diabetes, non-alcoholic fatty liver disease (NAFLD) and kidney disease, it is crucial to recognize that individual responses to the KD can vary significantly [102,103].

8. The Use of the KD in ND

Since the 1920s, the KD has been employed as a therapeutic approach for various neurological disorders, most notably in the management of drug-resistant refractory childhood epilepsy [92,104,105]. More recently, its potential role has been explored in the context of several neurodegenerative diseases, including AD [66] and PD [105,106], as well as in psychiatric diseases such as schizophrenia [107,108], depression [109,110] and Borderline Syndrome [111]. The success behind the use of the KD is linked to its action in common features shared by most CNS diseases, such as glucose hypometabolism, energetic deficits, imbalanced GABA and glutamate transmission, inflammation and oxidative stress [70,92]. Notably in some neurodegenerative diseases, patients present failures in the expression of glucose receptors GLUT1 and GLUT3, which may lead to glucose hypometabolism, contributing to the aggravation of diseases [92,112,113]. For instance, Glut1 deficiency syndrome (Glut1DS) causes a delay in glucose transport through the blood–brain barrier (BBB), leading to decreased cerebrospinal fluid glucose levels even in the presence of normal blood glucose [114]. In this context, the KD can enhance the provision of ketones at the expense of carbohydrates, potentially leading to an increased production of ATP, protecting neurons from energetic deficits. This is supported by a study that demonstrated a 22% increase in the expression of BOHb transporters, a key component of the KD, in patients with schizophrenia [113]. Moreover, the imbalance between GABA and glutamate represents a common feature contributing to the characteristic seizures observed in many neuropsychiatric disorders [70]. Notably, the Medium-Chain Triglyceride Ketogenic Diet (MCT KD) achieves seizure control via decanoic acid, which selectively inhibits AMPA receptors as demonstrated in preclinical models [31,92]. Furthermore, the anti-inflammatory properties of the KD are associated with its ability to shift microglial cells from a pro-inflammatory state to an anti-inflammatory state, offering promise in the management of mental illnesses [115]. Additionally, the KD, primarily via its ketone constituents, exerts control over oxidative stress by influencing various metabolic and signaling factors [70,92]. This multifaceted action underscores the broad spectrum of the KD’s therapeutic potential in treating various CNS diseases, thereby encouraging further exploration of its utility in conditions such as AD.

9. The Rationality of KD Use in AD

The progressive deposition of Aβ peptides and increased levels of hyperphosphorylated tau protein trigger neurodegeneration and impaired glucose metabolism [116]. Indeed, deficiencies of the GLUT1 receptor have been reported in AD, leading to impairment of glucose transport through the blood–brain barrier (BBB), causing hypometabolism of glucose in the CNS and consequently, starvation of neurons because of inefficient glycolysis [92,112,117]. Thus, an adjuvant therapy route to overcome the energy inefficiency of glucose in AD may act as an alternative fuel for brain metabolism [118,119]. In this case, the KD offers high levels of fat that, when metabolized in the liver, generate ketones which can supply the brain energetically and prevent neuronal death and synapse loss [41]. KD treatment may lead to neuroprotective effects by reducing Aβ damage to mitochondria, increasing ATP production, and reducing oxidative stress and glutamate toxicity [120]. Improvements in mitochondrial function have been attributed to biochemical changes resulting from the inhibition of glycolysis and increased KB formation [41]. Furthermore, KD treatment elevates ketone levels, and this increase exerts a neuroprotective effect on aging brain cells by diminishing the expression of inflammatory and apoptotic mediators [41,121,122].
Indeed, intriguing hypotheses propose a multifaceted impact of ketones on gene expression alterations and the modulation of cell signaling cascades. These processes, in turn, appear to regulate neuronal excitability, bolster antioxidant defenses and maintain the redox balance within cells [123,124]. In addition, there are suggestions that KD treatment may impact the deposition of Aβ or tau, slowing down the underlying disease process [92,125]. Moreover, the KD may act to reestablish the misbalance between the GABAergic and glutamatergic neurotransmission present in AD [126,127]. It has been reported that Aβ may increase AMPA currents, leading to glutamatergic hyperactivity, neurotoxicity and memory loss in AD [128]. In addition, the KD showed improvement in neuronal survival via the inhibition of AMPA receptors [31,92]. Another crucial aspect is that KD therapy contributes to the preservation of synaptic activity. This is achieved by enhancing the GABA/Glutamate ratio via elevated levels of Krebs cycle intermediates and by activating ATP-sensitive potassium channels via mitochondrial metabolism [129]. Hence, the implementation of the KD may attenuate neuronal hyperexcitability that has been described in preclinical and early clinical stages of AD [130,131].
Also, neuroinflammation, brain cell atrophy and apoptosis observed in AD may be related to the deleterious cellular effects of phosphate toxicity. It seems that hyperphosphorylated tau engages in the significant utilization of phosphate; consequently, a diminished dietary phosphorus intake emerges as a potential strategy to mitigate the risk of AD [46]. Interestingly, some studies suggest that the KD may increase the excretion of phosphorus in urine. This could potentially lead to lower phosphate levels in the body [46].

10. KD Results in Preclinical Models of AD

Studies with preclinical models demonstrate that the KD can provide benefits regarding inflammation [130,132,133], pain [134,135], cancer [136], metabolic disorders [137,138,139], epilepsy [140,141], aging [142], memory [143,144], mood [145,146] and neurodegenerative diseases [147,148,149]. Within this latter group, the benefits of the KD in preclinical models of AD involve changes at molecular and cellular levels, culminating in symptom improvement [125,130] (Table 2).
One potential mechanism by which the KD operates in mouse models of AD is via the direct reduction of Aβ. An intriguing study utilizing the APP mouse model (APP/V717I) revealed that after 43 days of KD treatment, there was an approximate 25% decrease in brain Aβ levels (Aβ40 and Aβ42) [131]. Nevertheless, it is worth noting that in this model, no significant memory improvement was observed when assessed using the object recognition test [131]. In a preclinical study, 5XFAD mice, exposed to a KD for an extended duration of 4 months, exhibited notable cognitive improvements alongside reduced neurotoxicity and diminished neuroinflammation [130]. Moreover, a significant decrease in the accumulation of Aβ plaques was observed in the immunohistochemical analysis of the hippocampus of treated KD mice [130]. This is an important result as the accumulation of Aβ favors neurotoxic events in the brain [150,151]. Of note, following an evaluation of spatial learning, mice subjected to a KD displayed enhanced efficiency with shorter latency times in comparison to standard-diet-treated mice. Moreover, apart from the behavioral improvements, the KD-treated mice also exhibited significant preservation of synaptic density, along with reductions in neuroinflammation and microgliosis [130].
Curiously, some studies have demonstrated an improvement in AD phenotype without changes in the deposition of Aβ in the brain. An investigation employing different AD models, including Aβ deposition (APP/PS1) and a tau protein accumulation model (Tg4510), underscores the advantageous effects of the KD [125]. In this study, five-month-old animals were subjected to a KD for duration of 3 months. Notably, upon analyzing the deposition of tau and Aβ in the cortical and hippocampal brain regions, no discernible differences were observed between the KD-treated mice and those on a control diet [125]. However, KD-treated mice exhibited improved motor performance in the rotarod test [125].
Another interesting fact about the KD is the direct effect on neuroinflammation, a hallmark of AD [152,153]. The anti-inflammatory effect of the KD in preclinical models has been studied [134,135,154]. This potential anti-inflammatory effect may be linked to a direct action on the mTOR pathway [155,156], PPARγ [157,158], NLRP3 inflammasome [159], in addition to other targets such as ketone bodies [160], adenosine [141,161] and the intestinal microbiota [162]. Mice fed with a KD for duration of 4 weeks exhibited a notable reduction in neuroinflammation induced by the administration of kainic acid (KA). This suggests that the diet’s anti-inflammatory effects are closely linked to the activation of PPARγ activation [133]. Consequently, the KD, via PPARγ activation, led to a decrease in TNF-α and NF-kB levels in the hippocampus following KA administration. Furthermore, the KD effectively inhibited the expression of cyclooxygenase (COX-2) and microsomal prostaglandin E2 synthase-1 (mPGES-1) in the hippocampus. These findings suggest that the KD can attenuate neuroinflammation by suppressing the COX-2 dependent pathway via PPARγ activation [133]. It is worth noting that the literature indicates that increasing PPARγ expression in certain brain regions may reduce and even prevent the neuroinflammatory response [133,163]. This is attributed to the pivotal role of PPARγ in regulating pro-inflammatory cytokines [164], modulating microglia [165], regulating macrophage activation [166] and inhibiting transcription factors [167].
In obese rats subjected to a KD, a notable reduction in the expression of APP and apoE mRNA has been observed [168]. Furthermore, KD treatment resulted in weight loss, lowered cholesterol and glucose levels, and improved insulin sensitivity, alongside a reduction in inflammatory markers. These findings may provide insight into the KD’s favorable effects on APP expression in the brain [168]. Additionally, aside from impacting Aβ and tau deposition and neuroinflammation, the KD appears to induce morphological changes in microglial and astroglial cells [169].
Another noteworthy factor is the impact of the KD on the intestinal microbiota [73,170,171]. Notably, in a study involving young healthy mice (12–14 weeks old) subjected to chronic KD treatment, several significant effects were observed. These included an increase in peripheral plasma ketone levels, which corresponded with enhanced diversity and the presence of beneficial gut microbiota. At the central level, the study revealed an augmentation in cerebral blood flow and alterations in the expression of various proteins, including mTOR, P-glycoprotein transports and eNOS [172]. These results demonstrate an improvement in neurovascular function resulting from the decrease in mTOR activity and eNOS activation [172]. These findings suggest that changes in the microbiota may lead to protein alterations that enhance neurovascular activity, thereby mitigating the risk of AD [172].
Apart from the classic or standard KD, there is evidence that various KD variations can also yield positive results against AD in animal models [173,174]. In one study, the authors examined the effects of an 8-week treatment with medium-chain fatty acids derived from MCTs, namely MCT8, MCT10 or supplementation with sunflower oil (5% by weight of the chow diet) in 21-month-old Wistar rats. The 8-week MCT diets exhibited cognitive and physiological improvements in the rats. MCT treatment led to an increase in the expression of growth factors in the hippocampus and promoted alterations in synaptic markers, transcription factors, protein synthesis and plasticity [174]. In addition, treatment with MCT10, but not MCT8, significantly improved novel object recognition memory as compared to a control diet, while social recognition was increased in both MCT groups [174]. Also, treatment with supplements capable of inducing ketogenesis has shown promising results [120,175,176]. Ketogenic agents are an alternative to the use of a diet capable of producing the ketone bodies necessary for energy metabolism in cells [177]. Some exogenous agents, such as 2-deoxy-D-glucose (2-DG) and the ester of D-β-hydroxybutyrate and R-1,3 butane diol, referred to as ketone ester (KE), demonstrate a beneficial effect in a model of triple-transgenic AD mice (3xTgAD) [120,175,176]. 2-deoxy-D-glucose (2-DG), when administered over a 7-week period to 6-month-old 3xTgAD mice, led to an elevation in serum ketone body levels, subsequently resulting in a reduction in the pathological effects of AD in these mice [120]. The application of 2-deoxy-D-glucose (2-DG) was associated with a decrease in Aβ load, an improvement in mitochondrial bioenergetic capacity and an increase in the expression of neurotrophic growth factors. Conversely, the ketogenic ester (KE), when incorporated into the mice’s diet, contributed to increased ketone body levels, consequently diminishing protein and lipid oxidation and enhancing ATP hydrolysis energy [175]. A reduction in Aβ and hyperphosphorylated tau deposition levels was also noted. Surprisingly, these effects appeared to influence the behavior of these animals. The 3xTgAD mice exhibited decreased anxious behavior and performed better on tests of learning and memory [176]. These results underscore the positive effects resulting from the induction of ketone bodies via exogenous ketogenic agents.
Table 2. KD results in preclinical models of AD.
Table 2. KD results in preclinical models of AD.
AuthorSpecies StudiedExperimental ModelDiet/
Supplementation
DurationEffects
Yao, J. et al. [120]Mice (♀)3xTgAD2-deoxy-D-glucose (2-DG)7 weeksReduced β-amyloid generation, increased β-amyloid clearance.
Enhanced mitochondrial bioenergetic capacity and increased the expression of neurotrophic growth factors.
Brownlow, M. L. et al. [125]MiceAPP/PS1-Tg4510Medium-chain triglyceride KD16 weeksImproved motor performance in rotarod test.
Xu, Y. et al. [130]Mice (♂)5XFADClassic KD4 monthsImproved spatial learning, spatial memory and working memory.
Restored number of neurons and synapses.
Reduced neuroinflammation, amyloid plaque deposition and microglial activation.
Van der Auwera, I. et al. [131]Mice (♀)APP/V717IClassic KD38 daysReduced levels of Aβ in brain tissue.
Liu, H. et al. [155]Rats (♂)Sprague-DawleyKD with or without medium-chain fatty acids30 daysEffects on the mTOR pathway and anti-inflammation action.
Mohamed, H. E. et al. [168]Rats (♂)Obesity induced with HFDClassic KD6 weeksImprovement of brain oxidative stress responses.
Downregulation of brain amyloid protein precursor, apolipoprotein E and caspase-3 mRNA expression.
Gzielo, K. et al. [169].Rats (♂)WistarClassic KD4 monthsMorphologically changes in microglial and astroglial cells.
Ma, D. et al. [172]Mice (♂)C57Bl/6Classic KD16 weeksIncreased CBF and P-glycoprotein transports on BBB.
Reduced mTOR and increased eNOS protein expressions.
Enhanced neurovascular functions.
Increased the abundance of beneficial gut microbiota.
Wang, D. & Mitchell, E. S. [174]Rats (♂)WistarMedium-chain triglyceride KD8 weeksIncreased expression of growth factors, alteration of synaptic markers, transcription factor, protein synthesis and plasticity.
Cognitive improvement and difference performance in object and social recognition tests.
Pawlosky, R. J. et al. [175]Mice (♂)3xTgADKetone ester (KE)8 monthsCorrected energy deficiencies in the hippocampus, improved biomarkers and reduced oxidative damage.
Kashiwaya, Y. et al. [176]Mice (♂)3xTgADKetone ester (KE)8 monthsImproved behavioral cognitive function and decreased.
Aβ and pTau pathologic changes.
Abbreviations: BBB = blood–brain barrier; CBF = cerebral blood flow; eNos = endothelial nitric oxide synthase. ♀ female and ♂ male.

11. KD Results in Clinical Trials

In addition to data acquired from preclinical research, several clinical studies demonstrate the feasibility and effectiveness of the KD in patients diagnosed with AD [94,178,179] (Table 3). In a randomized trial aimed at investigating the impact of a KD on patients with consistent AD diagnoses, those treated with a 12-week KD reported an improved quality of life and enhanced daily functionality [180]. Furthermore, the diet was found to be safe, with only mild adverse effects. Notably, it was observed that high levels of dietary lipids did not appear to pose significant cardiovascular risks for the patients [180]. It is worth mentioning that the sample size and duration of this trial are limiting factors. Specifically, a modified MKD, when provided to AD patients, is shown to enhance cognitive status while positively impacting metabolic and biomarker parameters [181]. Comparative analysis between the modified KD and a control group revealed improvements in cerebrospinal fluid biomarkers, including an increase in Aβ42 and a decrease in tau levels [181]. Other favorable outcomes encompassed increased uptake of brain ketone bodies, enhanced brain perfusion and improved peripheral lipid and glucose metabolism. Cognitive performance, as evaluated using the Free and Cued Selective Reminding Test (FCSRT), also exhibited notable improvements among patients [181].
Furthermore, it is important to note that studies have demonstrated the Mediterranean–KD’s modulating effect on the gut microbiome, both in patients with mild cognitive impairment [182] and in preclinical models [182]. A randomized, double-blind, cross-over pilot study conducted at a single center involved 17 subjects of whom 11 had mild cognitive impairment, while the remaining 6 displayed normal cognitive function. The study encompassed a 6-week intervention of the Modified Mediterranean–Ketogenic Diet (MMKD) and the American Heart Association Diet (AHAD) separated by 6-week washout periods. Assessments were made for gut microbiome composition, fecal short-chain fatty acids (SCFAs) and AD markers in cerebrospinal fluid (CSF), which included Aβ-40, Aβ-42, total tau and phosphorylated tau-181 (tau-p181), both before and after the dietary interventions. Interesting, the results of this study indicate that distinct gut microbial signatures could potentially serve as indicators of mild cognitive impairment, and further reveal that the MMKD can effectively reshape the gut microbiome and associated metabolites, correlating with improvements in AD biomarkers in cerebrospinal fluid (CSF).
The MCT diet applied to patients with AD also shows positive data regarding the pathological conditions of the disease [93,183,184]. Patients with mild to moderate AD sequentially ingested two distinct MCT supplements, each for a duration of one month: a blend of caprylic acid (55%) and capric acid (35%) (n = 11), followed by a wash-out period, and subsequently tricaprylin (95%; n = 6). Using PET imaging, the study quantified brain ketone (11C-acetoacetate) and glucose (FDG) uptake before and after each MCT intervention. As a result, the dietary treatment significantly enhanced overall brain energy metabolism in individuals with mild to moderate AD, while not impacting brain glucose utilization [93].
The MCT diet was also shown to be effective in producing ketosis and patients with AD demonstrated an overall cognitive improvement [183]. An interesting pilot study showed the KD Retention and feasibility trial involved a 3-month medium-chain triglyceride (MCT)-supplemented KD, followed by a 1-month washout period, among participants with clinical dementia ratings (CDR) of 0.5, 1 and 2. Throughout the study, data on urine acetoacetate, serum β-hydroxybutyrate, food intake records and safety were collected. The primary outcome of the study revealed that participants experienced an average improvement of 4.1 points in their scores on the Alzheimer’s Disease Assessment Scale-cognitive subscale (ADAS-Cog) from the beginning of the dietary intervention to its conclusion. However, these improvements on the ADAS-Cog waned after a one-month diet washout period [183].
Remarkably, noteworthy improvements were observed in short-term memory and processing speed among patients who underwent MCT diet treatment for several weeks [184]. A study involved 20 Japanese patients with mild-to-moderate AD, comprising eleven males and nine females. On separate days, these patients underwent neurocognitive tests 120 min after consuming 50 g of a ketogenic formula known as Ketonformula®, which contained 20 g of medium-chain triglycerides (MCTs). On alternate days, they consumed an isocaloric placebo formula lacking MCTs. Subsequently, the patients consumed ketogenic formula daily for up to 12 weeks and underwent monthly neurocognitive assessments. In the initial trial, following a single intake of the ketogenic formula, the patients exhibited a successful elevation in plasma levels of ketone bodies. However, there were no significant differences in cognitive test results between the administration of the ketogenic and placebo formulas. In the subsequent chronic intake trial of the ketogenic formula, 16 out of the 20 patients completed the 12-week regimen. After 8 weeks, patients showed significant improvement in their immediate and delayed logical memory test scores compared to their baseline scores. By the end of the 12-week period, they demonstrated significant improvements in the digit-symbol coding test and immediate logical memory test scores relative to their baseline [184].
A randomized placebo-controlled trial included six participants with mild cognitive impairment (MCI). This study evaluated the effect of the daily consumption of oil composed of medium-chain triglycerides (MCTs) for 24 weeks on serum ketone body concentrations, BOHb, apolipoprotein-E4 status and cognitive performance assessed using the Alzheimer’s disease Assessment Scale-Cognitive subscale (ADAS-Cog). The intake of MCT oil resulted in elevated concentrations of BOHb and demonstrated an enhancement in memory among subjects with MCI. It is important to note that the study had limitations, particularly the small sample size [185].
Another factor that gives more reliability and credibility to the use of the KD in the treatment of AD is the use of ketogenic agents in patients with mild to moderate conditions [186]. An oral ketogenic compound, AC-1202, was tested in subjects with probable AD to examine if ketosis could improve cognitive performance. AC-1202 rapidly elevated serum ketone bodies in AD patients and resulted in significant differences in ADAS-Cog scores compared to the placebo. The most prominent effects were observed in subjects without the APOE4 allele who adhered to the prescribed dosage regimen [187]. Furthermore, the daily consumption of caprylidene, a ketogenic compound that, upon metabolism, generates the ketones BOHb and acetoacetate, known for their ability to cross the BBB, resulted in increased blood flow within distinct brain regions among patients who lacked an apolipoprotein ε4 allele over a period of 45 days [188]. Moreover, the supplementation of MCT-based beverages has demonstrated cognitive enhancements in these patients [189,190]. Specifically, a daily intake of 30 g of ketogenic medium-chain triglycerides (kMCTs) over a span of 6 months generated sufficient ketones to notably improve brain energy levels in individuals with MCI. Notably, various aspects of cognitive function exhibited improvement in direct correlation with the enhanced brain energy levels achieved via kMCT supplementation. Future investigations will determine whether a larger sample size can validate the cognitive benefits of this kMCT dosage or if a higher dose is necessary. Nonetheless, the study’s authors underscored the feasibility of conducting long-term clinical trials with kMCTs and energy-equivalent placebos among older individuals. Subsequently, further research aimed at postponing age-related cognitive decline via the optimization of brain energy support with ketones is strongly justified [190].

12. Conclusions

This review aims to shed light on both preclinical and clinical studies exploring the potential of KD as an adjuvant therapy for AD. Given the absence of specific drugs or direct treatments targeting this neurological disorder, the most promising path for enhancing the quality of life for patients involves a synergistic approach, combining pharmaceutical interventions with complementary non-pharmacological strategies. Considering the energy metabolism deficits in neurons associated with AD, it appears reasonable to explore alternative approaches in combination with pharmacological treatments. By employing these complementary strategies alongside medications, we can potentially create a more resilient response within the body, thereby enhancing its ability to resist the progression of this disorder. Based on the analysis conducted in this study, it is recommended to conduct additional investigations within both preclinical and clinical models. This is essential to gain a deeper understanding of the effects of KD and to obtain more comprehensive data regarding the mechanisms via which this diet operates. While some animal models, such as those with APP mutations, have yielded promising results, a more comprehensive assessment encompassing various types of KD and diverse rodent models of Alzheimer’s disease is crucial. This comprehensive approach is vital to identify and elucidate the primary cellular pathways responsible for the beneficial effects of the diet in animals with AD pathology. For this, the use of knockout animals or those with the overexpression of key proteins linked to neuroprotective effects in the brain as PPARγ and NLRP3 can provide more concrete results. Simultaneously, there is a persistent requirement to enhance clinical studies by involving extensive patient cohorts, implementing more frequent follow-ups, refining cognitive assessment tests and refining sample collection methodologies for molecular analyses. Moreover, it is essential to conduct long-term studies to ascertain whether the favorable outcomes associated with the adoption of this diet can be consistently maintained over an extended period. In alignment with this, KD emerges as a promising adjuvant therapy option, although further clarification of its precise mechanism of action and potential side effects remains necessary. This study has lightened potential benefits of KD across both preclinical models and clinical research. Overall, the findings suggest that KD holds intriguing prospects for further enhancement and utilization in AD management. The different variations of this diet, along with dietary supplements inspired by its principles, may offer novel avenues for facing the progression of AD, potentially leading to innovative strategies for intervention.

Author Contributions

L.B.V. and T.P.D.O. designed the topics of this review. All authors contributed to the preparation and writing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Morais A.L. is a PIBIC FAPEMIG fellow. The neuropharmacology laboratory is receiving funds from Congresswoman Beatriz Cerqueira (PT-MG) directly from the House of Representatives (ALMG), Covenant No. 2071001094/2020.

Acknowledgments

All authors would like to thank T.P.D. for his expert support on the figure design. L.B.V. would like to thank Andrea Laura Prates for suggesting this thematic review issue. Additional information: The Graphical abstract and Figure 1 were drawn using pictures from Servier Medical Art. Servier Medical Art by Servier is licensed under a Creative Commons Attribution 3.0 Unported License (https://creativecommons.org/licenses/by/3.0/, accessed on 1 January 2021).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Takada, L.T.; Caramelli, P.; Radanovic, M.; Anghinah, R.; Hartmann, A.P.; Guariglia, C.C.; Bahia, V.S.; Nitrini, R. Prevalence of potentially reversible dementias in a dementia outpatient clinic of a tertiary university-affiliated hospital in Brazil. Arq. Neuropsiquiatr. 2003, 61, 925–929. [Google Scholar] [CrossRef] [PubMed]
  2. Reitz, C.; Brayne, C.; Mayeux, R. Epidemiology of Alzheimer disease. Nat. Rev. Neurol. 2011, 7, 137–152. [Google Scholar] [CrossRef] [PubMed]
  3. Ferri, C.P.; Prince, M.; Brayne, C.; Brodaty, H.; Fratiglioni, L.; Ganguli, M.; Hall, K.; Hasegawa, K.; Hendrie, H.; Huang, Y.; et al. Alzheimer’s Disease International. Global prevalence of dementia: A Delphi consensus study. Lancet 2005, 366, 2112–2117. [Google Scholar] [CrossRef] [PubMed]
  4. Schneider, J.A.; Arvanitakis, Z.; Leurgans, S.E.; Bennett, D.A. The neuropathology of probable Alzheimer disease and mild cognitive impairment. Ann. Neurol. 2009, 66, 200–208. [Google Scholar] [CrossRef] [PubMed]
  5. Qiu, C.; Kivipelto, M.; Von Strauss, E. Epidemiology of Alzheimer’s disease: Occurrence, determinants, and strategies toward intervention. Dialogues Clin. Neurosci. 2009, 11, 111–128. [Google Scholar] [CrossRef] [PubMed]
  6. Gutierrez, B.A.; Silva, H.S.; Guimarães, C.; Campino, A.C. Impacto econômico da doença de Alzheimer no Brasil: É possível melhorar a assistência e reduzir custos? Economic impact of Alzheimer’s Disease in Brazil: Is it possible to improve care and minimize costs? Cien. Saude Colet. 2014, 19, 4479–4486. [Google Scholar] [CrossRef] [PubMed]
  7. Knopman, D.S.; Amieva, H.; Petersen, R.C.; Chételat, G.; Holtzman, D.M.; Hyman, B.T.; Nixon, R.A.; Jones, D.T. Alzheimer disease. Nat. Rev. Dis. Primers 2021, 7, 33. [Google Scholar] [CrossRef]
  8. Corbett, A.; Pickett, J.; Burns, A.; Corcoran, J.; Dunnett, S.B.; Edison, P.; Hagan, J.J.; Holmes, C.; Jones, E.; Katona, C.; et al. Drug repositioning for Alzheimer’s disease. Nat. Rev. Drug Discov. 2012, 11, 833–846. [Google Scholar] [CrossRef]
  9. Stefanacci, R.G. The costs of Alzheimer’s disease and the value of effective therapies. Am. J. Manag. Care 2011, 17, S356–S362. [Google Scholar]
  10. Wong, W. Economic burden of Alzheimer disease and managed care considerations. Am. J. Manag. Care. 2020, 26, S177–S183. [Google Scholar]
  11. Vogt, A.S.; Jennings, G.T.; Mohsen, M.O.; Vogel, M.; Bachmann, M.F. Alzheimer’s Disease: A Brief History of Immunotherapies Targeting Amyloid β. Int. J. Mol. Sci. 2023, 24, 3895. [Google Scholar] [CrossRef]
  12. Rao, A.K.; Chou, A.; Bursley, B.; Smulofsky, J.; Jezequel, J. Systematic review of the effects of exercise on activities of daily living in people with Alzheimer’s disease. Am. J. Occup. Ther. 2014, 68, 50–56. [Google Scholar] [CrossRef] [PubMed]
  13. Hebert, L.E.; Weuve, J.; Scherr, P.A.; Evans, D.A. Alzheimer disease in the United States (2010–2050) estimated using the 2010 census. Neurology 2013, 80, 1778–1783. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, H.Y.; Panegyres, P.K. The Role of Ethnicity in Alzheimer’s Disease: Findings From The C-PATH Online Data Repository. J. Alzheimers Dis. 2016, 51, 515–523. [Google Scholar] [CrossRef]
  15. Viña, J.; Lloret, A. Why women have more Alzheimer’s disease than men: Gender and mitochondrial toxicity of amyloid-beta peptide. J. Alzheimers Dis. 2010, 20, S527–S533. [Google Scholar] [CrossRef] [PubMed]
  16. Farrer, L.A.; Cupples, L.A.; Haines, J.L.; Hyman, B.; Kukull, W.A.; Mayeux, R.; Myers, R.H.; Pericak-Vance, M.A.; Risch, N.; Van Duijn, C.M. Effects of age, sex, and ethnicity on the association between apolipoprotein E genotype and Alzheimer disease. A meta-analysis. APOE and Alzheimer Disease Meta Analysis Consortium. JAMA 1997, 278, 1349–1356. [Google Scholar] [CrossRef] [PubMed]
  17. Canevelli, M.; Quarata, F.; Remiddi, F.; Lucchini, F.; Lacorte, E.; Vanacore, N.; Bruno, G.; Cesari, M. Sex and gender differences in the treatment of Alzheimer’s disease: A systematic review of randomized controlled trials. Pharmacol. Res. 2017, 115, 218–223. [Google Scholar] [CrossRef] [PubMed]
  18. Selkoe, D.J. The cell biology of beta-amyloid precursor protein and presenilin in Alzheimer’s disease. Trends. Cell Biol. 1998, 8, 447–453. [Google Scholar] [CrossRef]
  19. Spillantini, M.G.; Goedert, M. Tau protein pathology in neurodegenerative diseases. Trends Neurosci. 1998, 21, 428–433. [Google Scholar] [CrossRef]
  20. Selkoe, D.J. Alzheimer’s disease: Genes, proteins, and therapy. Physiol. Rev. 2001, 81, 741–766. [Google Scholar] [CrossRef]
  21. Hardy, J.; Selkoe, D.J. The amyloid hypothesis of Alzheimer’s disease: Progress and problems on the road to therapeutics. Science 2002, 297, 353–356. [Google Scholar] [CrossRef] [PubMed]
  22. Turner, P.R.; O’Connor, K.; Tate, W.P.; Abraham, W.C. Roles of amyloid precursor protein and its fragments in regulating neural activity, plasticity and memory. Prog. Neurobiol. 2003, 70, 1–32. [Google Scholar] [CrossRef] [PubMed]
  23. Perez, R.G.; Zheng, H.; Van der Ploeg, L.H.; Koo, E.H. The beta-amyloid precursor protein of Alzheimer’s disease enhances neuron viability and modulates neuronal polarity. J. Neurosci. 1997, 17, 9407–9414. [Google Scholar] [CrossRef] [PubMed]
  24. Gauthier, S.; Wirth, Y.; Möbius, H.J. Effects of memantine on behavioural symptoms in Alzheimer’s disease patients: An analysis of the Neuropsychiatric Inventory (NPI) data of two randomised, controlled studies. Int. J Geriatr. Psychiatry 2005, 20, 459–464. [Google Scholar] [CrossRef] [PubMed]
  25. Tarawneh, R.; Holtzman, D.M. The clinical problem of symptomatic Alzheimer disease and mild cognitive impairment. Cold Spring Harb. Perspect. Med. 2012, 2, a006148. [Google Scholar] [CrossRef] [PubMed]
  26. Iqbal, K.; Wang, X.; Blanchard, J.; Liu, F.; Gong, C.X.; Grundke-Iqbal, I. Alzheimer’s disease neurofibrillary degeneration: Pivotal and multifactorial. Biochem. Soc. Trans. 2010, 38, 962–966. [Google Scholar] [CrossRef] [PubMed]
  27. de la Rubia Ortí, J.E.; Fernández, D.; Platero, F.; García-Pardo, M.P. Can Ketogenic Diet Improve Alzheimer’s Disease? Association With Anxiety, Depression, and Glutamate System. Front. Nutr. 2021, 8, 744398. [Google Scholar] [CrossRef]
  28. Herrmann, N.; Chau, S.A.; Kircanski, I.; Lanctôt, K.L. Current and emerging drug treatment options for Alzheimer’s disease: A systematic review. Drugs 2011, 71, 2031–2065. [Google Scholar] [CrossRef]
  29. Corey-Bloom, J. The ABC of Alzheimer’s disease: Cognitive changes and their management in Alzheimer’s disease and related dementias. Int. Psychogeriatr. 2002, 14, 51–75. [Google Scholar] [CrossRef]
  30. Haaksma, M.L.; Vilela, L.R.; Marengoni, A.; Calderón-Larrañaga, A.; Leoutsakos, J.S.; Olde Rikkert, M.G.M.; Melis, R.J.F. Comorbidity and progression of late onset Alzheimer’s disease: A systematic review. PLoS ONE 2017, 12, e0177044. [Google Scholar] [CrossRef]
  31. Pinto, A.; Bonucci, A.; Maggi, E.; Corsi, M.; Businaro, R. Anti-Oxidant and Anti-Inflammatory Activity of Ketogenic Diet: New Perspectives for Neuroprotection in Alzheimer’s Disease. Antioxidants 2018, 7, 63. [Google Scholar] [CrossRef] [PubMed]
  32. Verdile, G.; Keane, K.N.; Cruzat, V.F.; Medic, S.; Sabale, M.; Rowles, J.; Wijesekara, N.; Martins, R.N.; Fraser, P.E.; Newsholme, P. Inflammation and Oxidative Stress: The Molecular Connectivity between Insulin Resistance, Obesity, and Alzheimer’s Disease. Mediators Inflamm. 2015, 2015, 105828. [Google Scholar] [CrossRef] [PubMed]
  33. Tiwari, S.; Atluri, V.; Kaushik, A.; Yndart, A.; Nair, M. Alzheimer’s disease: Pathogenesis, diagnostics, and therapeutics. Int. J. Nanomed. 2019, 14, 5541–5554. [Google Scholar] [CrossRef] [PubMed]
  34. Ehret, M.J.; Chamberlin, K.W. Current Practices in the Treatment of Alzheimer Disease: Where is the Evidence After the Phase III Trials? Clin. Ther. 2015, 37, 1604–1616. [Google Scholar] [CrossRef]
  35. Farlow, M.R. Etiology and pathogenesis of Alzheimer’s disease. Am. J. Health Syst. Pharm. 1998, 55, S5–S10. [Google Scholar] [CrossRef] [PubMed]
  36. Ballard, C.; Gauthier, S.; Corbett, A.; Brayne, C.; Aarsland, D.; Jones, E. Alzheimer’s disease. Lancet 2011, 377, 1019–1031. [Google Scholar] [CrossRef] [PubMed]
  37. Scheltens, P.; Blennow, K.; Breteler, M.M.; de Strooper, B.; Frisoni, G.B.; Salloway, S.; van der Flier, W.M. Alzheimer’s disease. Lancet 2016, 388, 505–517. [Google Scholar] [CrossRef] [PubMed]
  38. Ahtiluoto, S.; Polvikoski, T.; Peltonen, M.; Solomon, A.; Tuomilehto, J.; Winblad, B.; Sulkava, R.; Kivipelto, M. Diabetes, Alzheimer disease, and vascular dementia: A population-based neuropathologic study. Neurology 2010, 75, 1195–1202. [Google Scholar] [CrossRef]
  39. Husain, M.A.; Laurent, B.; Plourde, M. APOE and Alzheimer’s Disease: From Lipid Transport to Physiopathology and Therapeutics. Front. Neurosci. 2021, 15, 630502. [Google Scholar] [CrossRef]
  40. Norton, S.; Matthews, F.E.; Barnes, D.E.; Yaffe, K.; Brayne, C. Potential for primary prevention of Alzheimer’s disease: An analysis of population-based data. Lancet Neurol. 2014, 13, 788–794. [Google Scholar] [CrossRef]
  41. Rusek, M.; Pluta, R.; Ułamek-Kozioł, M.; Czuczwar, S.J. Ketogenic Diet in Alzheimer’s Disease. Int. J. Mol. Sci. 2019, 20, 3892. [Google Scholar] [CrossRef] [PubMed]
  42. Hersi, M.; Irvine, B.; Gupta, P.; Gomes, J.; Birkett, N.; Krewski, D. Risk factors associated with the onset and progression of Alzheimer’s disease: A systematic review of the evidence. Neurotoxicology 2017, 61, 143–187. [Google Scholar] [CrossRef] [PubMed]
  43. Cheng, D.; Noble, J.; Tang, M.X.; Schupf, N.; Mayeux, R.; Luchsinger, J.A. Type 2 diabetes and late-onset Alzheimer’s disease. Dement. Geriatr. Cogn. Disord. 2011, 31, 424–430. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, L.; Xu, X.; Cao, H.; Li, H. Comparison of effects of different dietary interventions on cognitive function in Alzheimer’s disease: Protocol for systematic review and network meta-analysis. BMJ Open 2021, 11, e042997. [Google Scholar] [CrossRef]
  45. Lloret, A.; Monllor, P.; Esteve, D.; Cervera-Ferri, A.; Lloret, M.A. Obesity as a Risk Factor for Alzheimer’s Disease: Implication of Leptin and Glutamate. Front. Neurosci. 2019, 13, 508. [Google Scholar] [CrossRef] [PubMed]
  46. Brown, R.B. Stress, inflammation, depression, and dementia associated with phosphate toxicity. Mol. Biol. Rep. 2020, 47, 9921–9929. [Google Scholar] [CrossRef] [PubMed]
  47. Paradise, M.; Cooper, C.; Livingston, G. Systematic review of the effect of education on survival in Alzheimer’s disease. Int. Psychogeriatr. 2009, 21, 25–32. [Google Scholar] [CrossRef] [PubMed]
  48. Caamaño-Isorna, F.; Corral, M.; Montes-Martínez, A.; Takkouche, B. Education and dementia: A meta-analytic study. Neuroepidemiology 2006, 26, 226–232. [Google Scholar] [CrossRef]
  49. Zhang, T.; Liu, N.; Cao, H.; Wei, W.; Ma, L.; Li, H. Different Doses of Pharmacological Treatments for Mild to Moderate Alzheimer’s Disease: A Bayesian Network Meta-Analysis. Front. Pharmacol. 2020, 11, 778. [Google Scholar] [CrossRef]
  50. May, B.H.; Feng, M.; Hyde, A.J.; Hügel, H.; Chang, S.Y.; Dong, L.; Guo, X.; Zhang, A.L.; Lu, C.; Xue, C.C. Comparisons between traditional medicines and pharmacotherapies for Alzheimer disease: A systematic review and meta-analysis of cognitive outcomes. Int. J. Geriatr. Psychiatry 2018, 33, 449–458. [Google Scholar] [CrossRef]
  51. Shearer, J.; Green, C.; Ritchie, C.W.; Zajicek, J.P. Health state values for use in the economic evaluation of treatments for Alzheimer’s disease. Drugs Aging 2012, 29, 31–43. [Google Scholar] [CrossRef] [PubMed]
  52. Ulamek-Koziol, M.; Pluta, R. To treat or not to treat Alzheimer’s disease by the ketogenic diet? That is the question. Neural Regen. Res. 2020, 15, 857–858. [Google Scholar] [PubMed]
  53. Salek, S.S.; Walker, M.D.; Bayer, A.J. A review of quality of life in Alzheimer’s disease. Part 2: Issues in assessing drug effects. Pharmacoeconomics 1998, 14, 613–627. [Google Scholar] [CrossRef] [PubMed]
  54. Overshott, R.; Byrne, J.; Burns, A. Nonpharmacological and pharmacological interventions for symptoms in Alzheimer’s disease. Expert Rev. Neurother. 2004, 4, 809–821. [Google Scholar] [CrossRef] [PubMed]
  55. Sandoz, M.; Démonet, J.F.; Fossard, M. Theory of mind and cognitive processes in aging and Alzheimer type dementia: A systematic review. Aging Ment. Health 2014, 18, 815–827. [Google Scholar] [CrossRef] [PubMed]
  56. Massoud, F.; Léger, G.C. Pharmacological treatment of Alzheimer disease. Can. J. Psychiatry 2011, 56, 579–588. [Google Scholar] [CrossRef]
  57. Seibert, M.; Mühlbauer, V.; Holbrook, J.; Voigt-Radloff, S.; Brefka, S.; Dallmeier, D.; Denkinger, M.; Schönfeldt-Lecuona, C.; Klöppel, S.; Von Arnim, C.A.F. Efficacy and safety of pharmacotherapy for Alzheimer’s disease and for behavioural and psychological symptoms of dementia in older patients with moderate and severe functional impairments: A systematic review of controlled trials. Alzheimers Res. Ther. 2021, 13, 131. [Google Scholar] [CrossRef]
  58. Fink, H.A.; Linskens, E.J.; MacDonald, R.; Silverman, P.C.; McCarten, J.R.; Talley, K.M.C.; Forte, M.L.; Desai, P.J.; Nelson, V.A.; Miller, M.A.; et al. Benefits and Harms of Prescription Drugs and Supplements for Treatment of Clinical Alzheimer-Type Dementia. Ann. Intern. Med. 2020, 172, 656–668. [Google Scholar] [CrossRef]
  59. Hansen, R.A.; Gartlehner, G.; Lohr, K.N.; Kaufer, D.I. Functional outcomes of drug treatment in Alzheimer’s disease: A systematic review and meta-analysis. Drugs Aging 2007, 24, 155–167. [Google Scholar] [CrossRef]
  60. Mendiola-Precoma, J.; Berumen, L.C.; Padilla, K.; Garcia-Alcocer, G. Therapies for Prevention and Treatment of Alzheimer’s Disease. BioMed Res. Int. 2016, 2016, 2589276. [Google Scholar] [CrossRef]
  61. Evans, J.G.; Wilcock, G.; Birks, J. Evidence-based pharmacotherapy of Alzheimer’s disease. Int. J. Neuropsychopharmacol. 2004, 7, 351–369. [Google Scholar] [CrossRef] [PubMed]
  62. Wilcock, G.K.; Ballard, C.G.; Cooper, J.A.; Loft, H. Memantine for agitation/aggression and psychosis in moderately severe to severe Alzheimer’s disease: A pooled analysis of 3 studies. J. Clin. Psychiatry 2008, 69, 341–348. [Google Scholar] [CrossRef] [PubMed]
  63. Olazarán, J.; Reisberg, B.; Clare, L.; Cruz, I.; Peña-Casanova, J.; Del Ser, T.; Woods, B.; Beck, C.; Auer, S.; Lai, C.; et al. Nonpharmacological therapies in Alzheimer’s disease: A systematic review of efficacy. Dement. Geriatr. Cogn. Disord. 2010, 30, 161–178. [Google Scholar] [CrossRef] [PubMed]
  64. Ruthirakuhan, M.; Luedke, A.C.; Tam, A.; Goel, A.; Kurji, A.; Garcia, A. Use of physical and intellectual activities and socialization in the management of cognitive decline of aging and in dementia: A review. J. Aging Res. 2012, 2012, 384875. [Google Scholar] [CrossRef] [PubMed]
  65. Kang, M.J.; Kim, S.M.; Han, S.E.; Bae, J.H.; Yu, W.J.; Park, M.Y.; Ku, S.; Yang, Y. Effect of Paper-Based Cognitive Training in Early Stage of Alzheimer’s Dementia. Dement. Neurocogn. Disord. 2019, 18, 62–68. [Google Scholar] [CrossRef] [PubMed]
  66. Broom, G.M.; Shaw, I.C.; Rucklidge, J.J. The ketogenic diet as a potential treatment and prevention strategy for Alzheimer’s disease. Nutrition 2019, 60, 118–121. [Google Scholar] [CrossRef] [PubMed]
  67. Gasior, M.; Rogawski, M.A.; Hartman, A.L. Neuroprotective and disease-modifying effects of the ketogenic diet. Behav. Pharmacol. 2006, 17, 431–439. [Google Scholar] [CrossRef]
  68. Paoli, A.; Rubini, A.; Volek, J.S.; Grimaldi, K.A. Beyond weight loss: A review of the therapeutic uses of very-low-carbohydrate (ketogenic) diets. Eur. J. Clin. Nutr. 2013, 67, 789–796. [Google Scholar] [CrossRef]
  69. Zupec-Kania, B.A.; Spellman, E. An overview of the ketogenic diet for pediatric epilepsy. Nutr. Clin. Pract. 2008, 23, 589–596. [Google Scholar] [CrossRef]
  70. Norwitz, N.G.; Sethi, S.; Palmer, C.M. Ketogenic diet as a metabolic treatment for mental illness. Curr. Opin. Endocrinol. Diabetes Obes. 2020, 27, 269–274. [Google Scholar] [CrossRef]
  71. Włodarek, D. Role of Ketogenic Diets in Neurodegenerative Diseases (Alzheimer’s Disease and Parkinson’s Disease). Nutrients 2019, 11, 169. [Google Scholar] [CrossRef] [PubMed]
  72. Barry, D.; Ellul, S.; Watters, L.; Lee, D.; Haluska, R., Jr.; White, R. The ketogenic diet in disease and development. Int. J. Dev. Neurosci. 2018, 68, 53–58. [Google Scholar] [CrossRef] [PubMed]
  73. Park, S.; Zhang, T.; Wu, X.; Qiu, J.Y. Ketone production by ketogenic diet and by intermittent fasting has different effects on the gut microbiota and disease progression in an Alzheimer’s disease rat model. J. Clin. Biochem. Nutr. 2020, 67, 188–198. [Google Scholar] [CrossRef] [PubMed]
  74. Lange, K.W.; Lange, K.M.; Makulska-Gertruda, E.; Nakamura, Y.; Reissmann, A.; Kanaya, S.; Hauser, J. Ketogenic diets and Alzheimer’s disease. Food Sci. Hum. Wellness 2017, 6, 1–9. [Google Scholar] [CrossRef]
  75. Boison, D. New insights into the mechanisms of the ketogenic diet. Curr. Opin. Neurol. 2017, 30, 187–192. [Google Scholar] [CrossRef] [PubMed]
  76. Dhamija, R.; Eckert, S.; Wirrell, E. Ketogenic diet. Can. J. Neurol. Sci. 2013, 40, 158–167. [Google Scholar] [CrossRef] [PubMed]
  77. Silva, B.; Mantha, O.L.; Schor, J.; Pascual, A.; Plaçais, P.Y.; Pavlowsky, A.; Preat, T. Glia fuel neurons with locally synthesized ketone bodies to sustain memory under starvation. Nat. Metab. 2022, 4, 213–224. [Google Scholar] [CrossRef]
  78. Higashino-Matsui, Y.; Shirato, K.; Suzuki, Y.; Kawashima, Y.; Someya, Y.; Sato, S.; Shiraishi, A.; Jinde, M.; Matsumoto, A.; Ideno, H.; et al. Age-related effects of fasting on ketone body production during lipolysis in rats. Environ. Health Prev. Med. 2012, 17, 157–163. [Google Scholar] [CrossRef]
  79. Paoli, A.; Bosco, G.; Camporesi, E.M.; Mangar, D. Ketosis, ketogenic diet and food intake control: A complex relationship. Front. Psychol. 2015, 6, 27. [Google Scholar] [CrossRef]
  80. Evans, M.; Cogan, K.E.; Egan, B. Metabolism of ketone bodies during exercise and training: Physiological basis for exogenous supplementation. J. Physiol. 2017, 595, 2857–2871. [Google Scholar] [CrossRef]
  81. Qian, M.; Wu, N.; Li, L.; Yu, W.; Ouyang, H.; Liu, X.; He, Y.; Al-Mureish, A. Effect of Elevated Ketone Body on Maternal and Infant Outcome of Pregnant Women with Abnormal Glucose Metabolism During Pregnancy. Diabetes Metab. Syndr. Obes. 2020, 13, 4581–4588. [Google Scholar] [CrossRef] [PubMed]
  82. Newman, J.C.; Verdin, E. Ketone bodies as signaling metabolites. Trends Endocrinol. Metab. 2014, 25, 42–52. [Google Scholar] [CrossRef]
  83. Puchalska, P.; Crawford, P.A. Multi-dimensional Roles of Ketone Bodies in Fuel Metabolism, Signaling, and Therapeutics. Cell Metab. 2017, 25, 262–284. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, Y.H.; Liu, C.L.; Chiu, W.C.; Twu, Y.C.; Liao, Y.J. HMGCS2 Mediates Ketone Production and Regulates the Proliferation and Metastasis of Hepatocellular Carcinoma. Cancers 2019, 11, 1876. [Google Scholar] [CrossRef] [PubMed]
  85. Hegardt, F.G. Mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase: A control enzyme in ketogenesis. Biochem. J. 1999, 338, 569–582. [Google Scholar] [CrossRef] [PubMed]
  86. Orii, K.E.; Fukao, T.; Song, X.Q.; Mitchell, G.A.; Kondo, N. Liver-specific silencing of the human gene encoding succinyl-CoA: 3-ketoacid CoA transferase. Tohoku J. Exp. Med. 2008, 215, 227–236. [Google Scholar] [CrossRef] [PubMed]
  87. Martínez-Reyes, I.; Chandel, N.S. Mitochondrial TCA cycle metabolites control physiology and disease. Nat. Commun. 2020, 11, 102. [Google Scholar] [CrossRef]
  88. Roubergue, A.; Philibert, B.; Gautier, A.; Kuster, A.; Markowicz, K.; Billette de Villemeur, T.; Vuillaumier-Barrot, S.; Nicole, S.; Roze, E.; Doummar, D. Excellent response to a ketogenic diet in a patient with alternating hemiplegia of childhood. JIMD Rep. 2015, 15, 7–12. [Google Scholar]
  89. Kim, J.A.; Yoon, J.R.; Lee, E.J.; Lee, J.S.; Kim, J.T.; Kim, H.D.; Kang, H.C. Efficacy of the classic ketogenic and the modified Atkins diets in refractory childhood epilepsy. Epilepsia 2016, 57, 51–58. [Google Scholar] [CrossRef]
  90. Zhu, H.; Bi, D.; Zhang, Y.; Kong, C.; Du, J.; Wu, X.; Wei, Q.; Qin, H. Ketogenic diet for human diseases: The underlying mechanisms and potential for clinical implementations. Signal Transduct. Target Ther. 2022, 7, 11. [Google Scholar] [CrossRef]
  91. Liu, Y.M.; Wang, H.S. Medium-chain triglyceride ketogenic diet, an effective treatment for drug-resistant epilepsy and a comparison with other ketogenic diets. Biomed. J. 2013, 36, 9–15. [Google Scholar] [CrossRef] [PubMed]
  92. Augustin, K.; Khabbush, A.; Williams, S.; Eaton, S.; Orford, M.; Cross, J.H.; Heales, S.J.R.; Walker, M.C.; Williams, R.S.B. Mechanisms of action for the medium-chain triglyceride ketogenic diet in neurological and metabolic disorders. Lancet Neurol. 2018, 17, 84–93. [Google Scholar] [CrossRef] [PubMed]
  93. Croteau, E.; Castellano, C.A.; Richard, M.A.; Fortier, M.; Nugent, S.; Lepage, M.; Duchesne, S.; Whittingstall, K.; Turcotte, É.E.; Bocti, C.; et al. Ketogenic Medium Chain Triglycerides Increase Brain Energy Metabolism in Alzheimer’s Disease. J. Alzheimers Dis. 2018, 64, 551–561. [Google Scholar] [CrossRef] [PubMed]
  94. Caprio, M.; Infante, M.; Moriconi, E.; Armani, A.; Fabbri, A.; Mantovani, G.; Mariani, S.; Lubrano, C.; Poggiogalle, E.; Migliaccio, S.; et al. Very-low-calorie ketogenic diet (VLCKD) in the management of metabolic diseases: Systematic review and consensus statement from the Italian Society of Endocrinology (SIE). J. Endocrinol. Investig. 2019, 42, 1365–1386. [Google Scholar] [CrossRef] [PubMed]
  95. Youm, Y.H.; Nguyen, K.Y.; Grant, R.W.; Goldberg, E.L.; Bodogai, M.; Kim, D.; D’Agostino, D.; Planavsky, N.; Lupfer, C.; Kanneganti, T.D.; et al. The ketone metabolite β-hydroxybutyrate blocks NLRP3 inflammasome-mediated inflammatory disease. Nat. Med. 2015, 21, 263–269. [Google Scholar] [CrossRef] [PubMed]
  96. Greco, T.; Glenn, T.C.; Hovda, D.A.; Prins, M.L. Ketogenic diet decreases oxidative stress and improves mitochondrial respiratory complex activity. J. Cereb. Blood Flow Metab. 2016, 36, 1603–1613. [Google Scholar] [CrossRef]
  97. Dowis, K.; Banga, S. The Potential Health Benefits of the Ketogenic Diet: A Narrative Review. Nutrients 2021, 13, 1654. [Google Scholar] [CrossRef] [PubMed]
  98. Gupta, L.; Khandelwal, D.; Kalra, S.; Gupta, P.; Dutta, D.; Aggarwal, S. Ketogenic diet in endocrine disorders: Current perspectives. J. Postgrad. Med. 2017, 63, 242–251. [Google Scholar]
  99. Charlot, A.; Zoll, J. Beneficial Effects of the Ketogenic Diet in Metabolic Syndrome: A Systematic Review. Diabetology 2022, 3, 292–309. [Google Scholar] [CrossRef]
  100. Bostock, E.C.S.; Kirkby, K.C.; Taylor, B.V.; Hawrelak, J.A. Consumer Reports of “Keto Flu” Associated With the Ketogenic Diet. Front. Nutr. 2020, 7, 20. [Google Scholar] [CrossRef]
  101. Dashti, H.M.; Mathew, T.C.; Hussein, T.; Asfar, S.K.; Behbahani, A.; Khoursheed, M.A.; Al-Sayer, H.M.; Bo-Abbas, Y.Y.; Al-Zaid, N.S. Long-term effects of a ketogenic diet in obese patients. Exp. Clin. Cardiol. 2004, 9, 200–205. [Google Scholar] [PubMed]
  102. Batch, J.T.; Lamsal, S.P.; Adkins, M.; Sultan, S.; Ramirez, M.N. Advantages and Disadvantages of the Ketogenic Diet: A Review Article. Cureus 2020, 12, e9639. [Google Scholar] [CrossRef] [PubMed]
  103. Muscogiuri, G.; El Ghoch, M.; Colao, A.; Hassapidou, M.; Yumuk, V.; Busetto, L.; Obesity Management Task Force (OMTF) of the European Association for the Study of Obesity (EASO). European Guidelines for Obesity Management in Adults with a Very Low-Calorie Ketogenic Diet: A Systematic Review and Meta-Analysis. Obes. Facts 2021, 14, 222–245. [Google Scholar] [CrossRef] [PubMed]
  104. Martin, K.; Jackson, C.F.; Levy, R.G.; Cooper, P.N. Ketogenic diet and other dietary treatments for epilepsy. Cochrane Database Syst. Rev. 2016, 9, CD001903. [Google Scholar] [CrossRef] [PubMed]
  105. Liu, H.; Yang, Y.; Wang, Y.; Tang, H.; Zhang, F.; Zhang, Y.; Zhao, Y. Ketogenic diet for treatment of intractable epilepsy in adults: A meta-analysis of observational studies. Epilepsia Open 2018, 3, 9–17. [Google Scholar] [CrossRef]
  106. Vanitallie, T.B.; Nonas, C.; Di Rocco, A.; Boyar, K.; Hyams, K.; Heymsfield, S.B. Treatment of Parkinson disease with diet-induced hyperketonemia: A feasibility study. Neurology 2005, 64, 728–730. [Google Scholar] [CrossRef]
  107. Palmer, C.M.; Gilbert-Jaramillo, J.; Westman, E.C. The ketogenic diet and remission of psychotic symptoms in schizophrenia: Two case studies. Schizophr. Res. 2019, 208, 439–440. [Google Scholar] [CrossRef]
  108. Sarnyai, Z.; Kraeuter, A.K.; Palmer, C.M. Ketogenic diet for schizophrenia: Clinical implication. Curr. Opin. Psychiatry 2019, 32, 394–401. [Google Scholar] [CrossRef]
  109. Murphy, P.; Likhodii, S.; Nylen, K.; Burnham, W.M. The antidepressant properties of the ketogenic diet. Biol. Psychiatry 2004, 56, 981–983. [Google Scholar] [CrossRef]
  110. Włodarczyk, A.; Cubała, W.J.; Stawicki, M. Ketogenic diet for depression: A potential dietary regimen to maintain euthymia? Prog. Neuropsychopharmacol. Biol. Psychiatry 2021, 109, 110257. [Google Scholar] [CrossRef]
  111. Phelps, J.R.; Siemers, S.V.; El-Mallakh, R.S. The ketogenic diet for type II bipolar disorder. Neurocase 2013, 19, 423–426. [Google Scholar] [CrossRef] [PubMed]
  112. Kass, H.R.; Winesett, S.P.; Bessone, S.K.; Turner, Z.; Kossoff, E.H. Use of dietary therapies amongst patients with GLUT1 deficiency syndrome. Seizure 2016, 35, 83–87. [Google Scholar] [CrossRef] [PubMed]
  113. Sullivan, C.R.; Mielnik, C.A.; Funk, A.; O’Donovan, S.M.; Bentea, E.; Pletnikov, M.; Ramsey, A.J.; Wen, Z.; Rowland, L.M.; McCullumsmith, R.E. Measurement of lactate levels in postmortem brain, iPSCs, and animal models of schizophrenia. Sci. Rep. 2019, 9, 5087. [Google Scholar] [CrossRef] [PubMed]
  114. Rauchenzauner, M.; Klepper, J.; Leiendecker, B.; Luef, G.; Rostasy, K.; Ebenbichler, C. The ketogenic diet in children with Glut1 deficiency syndrome and epilepsy. J. Pediatr. 2008, 153, 716–718. [Google Scholar] [CrossRef]
  115. Morris, G.; Puri, B.K.; Maes, M.; Olive, L.; Berk, M.; Carvalho, A.F. The role of microglia in neuroprogressive disorders: Mechanisms and possible neurotherapeutic effects of induced ketosis. Prog. Neuropsychopharmacol. Biol. Psychiatry 2020, 99, 109858. [Google Scholar] [CrossRef] [PubMed]
  116. Castellano, C.A.; Nugent, S.; Paquet, N.; Tremblay, S.; Bocti, C.; Lacombe, G.; Imbeault, H.; Turcotte, É.; Fulop, T.; Cunnane, S.C. Lower brain 18F-fluorodeoxyglucose uptake but normal 11C-acetoacetate metabolism in mild Alzheimer’s disease dementia. J. Alzheimers Dis. 2015, 43, 1343–1353. [Google Scholar] [CrossRef] [PubMed]
  117. Winkler, E.A.; Nishida, Y.; Sagare, A.P.; Rege, S.V.; Bell, R.D.; Perlmutter, D.; Sengillo, J.D.; Hillman, S.; Kong, P.; Nelson, A.R.; et al. GLUT1 reductions exacerbate Alzheimer’s disease vasculo-neuronal dysfunction and degeneration. Nat. Neurosci. 2015, 18, 521–530. [Google Scholar] [CrossRef] [PubMed]
  118. Thelen, M.; Brown-Borg, H.M. Does Diet Have a Role in the Treatment of Alzheimer’s Disease? Front. Aging Neurosci. 2020, 12, 617071. [Google Scholar] [CrossRef]
  119. Yan, X.; Hu, Y.; Wang, B.; Wang, S.; Zhang, X. Metabolic Dysregulation Contributes to the Progression of Alzheimer’s Disease. Front. Neurosci. 2020, 14, 530219. [Google Scholar] [CrossRef]
  120. Yao, J.; Chen, S.; Mao, Z.; Cadenas, E.; Brinton, R.D. 2-Deoxy-D-glucose treatment induces ketogenesis, sustains mitochondrial function, and reduces pathology in female mouse model of Alzheimer’s disease. PLoS ONE 2011, 6, e21788. [Google Scholar] [CrossRef]
  121. Gough, S.M.; Casella, A.; Ortega, K.J.; Hackam, A.S. Neuroprotection by the Ketogenic Diet: Evidence and Controversies. Front. Nutr. 2021, 8, 782657. [Google Scholar] [CrossRef] [PubMed]
  122. Maalouf, M.; Rho, J.M.; Mattson, M.P. The neuroprotective properties of calorie restriction, the ketogenic diet, and ketone bodies. Brain Res. Rev. 2009, 59, 294–315. [Google Scholar] [CrossRef] [PubMed]
  123. García-Rodríguez, D.; Giménez-Cassina, A. Ketone Bodies in the Brain Beyond Fuel Metabolism: From Excitability to Gene Expression and Cell Signaling. Front. Mol. Neurosci. 2021, 14, 732120. [Google Scholar] [CrossRef] [PubMed]
  124. Henderson, S.T. Ketone bodies as a therapeutic for Alzheimer’s disease. Neurotherapeutics 2008, 5, 470–480. [Google Scholar] [CrossRef]
  125. Brownlow, M.L.; Benner, L.; D’Agostino, D.; Gordon, M.N.; Morgan, D. Ketogenic diet improves motor performance but not cognition in two mouse models of Alzheimer’s pathology. PLoS ONE 2013, 8, e75713. [Google Scholar] [CrossRef] [PubMed]
  126. Li, Y.; Sun, H.; Chen, Z.; Xu, H.; Bu, G.; Zheng, H. Implications of GABAergic Neurotransmission in Alzheimer’s Disease. Front. Aging Neurosci. 2016, 8, 31. [Google Scholar] [CrossRef] [PubMed]
  127. Wang, R.; Reddy, P.H. Role of Glutamate and NMDA Receptors in Alzheimer’s Disease. J. Alzheimers Dis. 2017, 57, 1041–1048. [Google Scholar] [CrossRef] [PubMed]
  128. Whitcomb, D.J.; Hogg, E.L.; Regan, P.; Piers, T.; Narayan, P.; Whitehead, G.; Winters, B.L.; Kim, D.H.; Kim, E.; St George-Hyslop, P.; et al. Intracellular oligomeric amyloid-beta rapidly regulates GluA1 subunit of AMPA receptor in the hippocampus. Sci. Rep. 2015, 5, 10934. [Google Scholar] [CrossRef]
  129. Kim, D.Y.; Abdelwahab, M.G.; Lee, S.H.; O’Neill, D.; Thompson, R.J.; Duff, H.J.; Sullivan, P.G.; Rho, J.M. Ketones prevent oxidative impairment of hippocampal synaptic integrity through KATP channels. PLoS ONE 2015, 10, e0119316. [Google Scholar] [CrossRef]
  130. Xu, Y.; Jiang, C.; Wu, J.; Liu, P.; Deng, X.; Zhang, Y.; Peng, B.; Zhu, Y. Ketogenic diet ameliorates cognitive impairment and neuroinflammation in a mouse model of Alzheimer’s disease. CNS Neurosci. Ther. 2022, 28, 580–592. [Google Scholar] [CrossRef]
  131. Van der Auwera, I.; Wera, S.; Van Leuven, F.; Henderson, S.T. A ketogenic diet reduces amyloid beta 40 and 42 in a mouse model of Alzheimer’s disease. Nutr. Metab. 2005, 2, 28. [Google Scholar] [CrossRef] [PubMed]
  132. Schmidt, T.; Harmon, D.M.; Kludtke, E.; Mickow, A.; Simha, V.; Kopecky, S. Dramatic elevation of LDL cholesterol from ketogenic-dieting: A Case Series. Am. J. Prev. Cardiol. 2023, 14, 100495. [Google Scholar] [CrossRef] [PubMed]
  133. Jeong, E.A.; Jeon, B.T.; Shin, H.J.; Kim, N.; Lee, D.H.; Kim, H.J.; Kang, S.S.; Cho, G.J.; Choi, W.S.; Roh, G.S. Ketogenic diet-induced peroxisome proliferator-activated receptor-γ activation decreases neuroinflammation in the mouse hippocampus after kainic acid-induced seizures. Exp. Neurol. 2011, 232, 195–202. [Google Scholar] [CrossRef] [PubMed]
  134. Ruskin, D.N.; Sturdevant, I.C.; Wyss, L.S.; Masino, S.A. Ketogenic diet effects on inflammatory allodynia and ongoing pain in rodents. Sci. Rep. 2021, 11, 725. [Google Scholar] [CrossRef] [PubMed]
  135. Ruskin, D.N.; Kawamura, M.; Masino, S.A. Reduced pain and inflammation in juvenile and adult rats fed a ketogenic diet. PLoS ONE 2009, 4, e8349. [Google Scholar] [CrossRef]
  136. Zou, Y.; Fineberg, S.; Pearlman, A.; Feinman, R.D.; Fine, E.J. The effect of a ketogenic diet and synergy with rapamycin in a mouse model of breast cancer. PLoS ONE 2020, 15, e0233662. [Google Scholar] [CrossRef]
  137. Cooper, M.A.; Menta, B.W.; Perez-Sanchez, C.; Jack, M.M.; Khan, Z.W.; Ryals, J.M.; Winter, M.; Wright, D.E. A ketogenic diet reduces metabolic syndrome-induced allodynia and promotes peripheral nerve growth in mice. Exp. Neurol. 2018, 306, 149–157. [Google Scholar] [CrossRef]
  138. Poplawski, M.M.; Mastaitis, J.W.; Isoda, F.; Grosjean, F.; Zheng, F.; Mobbs, C.V. Reversal of diabetic nephropathy by a ketogenic diet. PLoS ONE 2011, 6, e18604. [Google Scholar] [CrossRef]
  139. Okuda, T.; Morita, N. A very low carbohydrate ketogenic diet prevents the progression of hepatic steatosis caused by hyperglycemia in a juvenile obese mouse model. Nutr. Diabetes 2012, 2, e50. [Google Scholar] [CrossRef]
  140. Todorova, M.T.; Tandon, P.; Madore, R.A.; Stafstrom, C.E.; Seyfried, T.N. The ketogenic diet inhibits epileptogenesis in EL mice: A genetic model for idiopathic epilepsy. Epilepsia 2000, 41, 933–940. [Google Scholar] [CrossRef]
  141. Masino, S.A.; Li, T.; Theofilas, P.; Sandau, U.S.; Ruskin, D.N.; Fredholm, B.B.; Geiger, J.D.; Aronica, E.; Boison, D. A ketogenic diet suppresses seizures in mice through adenosine A1 receptors. J. Clin. Investig. 2011, 121, 2679–2683. [Google Scholar] [CrossRef] [PubMed]
  142. Roberts, M.N.; Wallace, M.A.; Tomilov, A.A.; Zhou, Z.; Marcotte, G.R.; Tran, D.; Perez, G.; Gutierrez-Casado, E.; Koike, S.; Knotts, T.A.; et al. A Ketogenic Diet Extends Longevity and Healthspan in Adult Mice. Cell Metab. 2017, 26, 539–546. [Google Scholar] [CrossRef] [PubMed]
  143. Newman, J.C.; Covarrubias, A.J.; Zhao, M.; Yu, X.; Gut, P.; Ng, C.P.; Huang, Y.; Haldar, S.; Verdin, E. Ketogenic Diet Reduces Midlife Mortality and Improves Memory in Aging Mice. Cell Metab. 2017, 26, 547–557. [Google Scholar] [CrossRef] [PubMed]
  144. Jiang, Y.; Lu, Y.; Jia, M.; Wang, X.; Zhang, Z.; Hou, Q.; Wang, B. Ketogenic diet attenuates spatial and item memory impairment in pentylenetetrazol-kindled rats. Brain Res. 2016, 1646, 451–458. [Google Scholar] [CrossRef] [PubMed]
  145. Guan, Y.F.; Huang, G.B.; Xu, M.D.; Gao, F.; Lin, S.; Huang, J.; Wang, J.; Li, Y.Q.; Wu, C.H.; Yao, S.; et al. Anti-depression effects of ketogenic diet are mediated via the restoration of microglial activation and neuronal excitability in the lateral habenula. Brain Behav. Immun. 2020, 88, 748–762. [Google Scholar] [CrossRef] [PubMed]
  146. Huang, J.; Li, Y.Q.; Wu, C.H.; Zhang, Y.L.; Zhao, S.T.; Chen, Y.J.; Deng, Y.H.; Xuan, A.; Sun, X.D. The effect of ketogenic diet on behaviors and synaptic functions of naive mice. Brain Behav. 2019, 9, e01246. [Google Scholar] [CrossRef] [PubMed]
  147. Ruskin, D.N.; Ross, J.L.; Kawamura, M., Jr.; Ruiz, T.L.; Geiger, J.D.; Masino, S.A. A ketogenic diet delays weight loss and does not impair working memory or motor function in the R6/2 1J mouse model of Huntington’s disease. Physiol. Behav. 2011, 103, 501–507. [Google Scholar] [CrossRef]
  148. Shaafi, S.; Najmi, S.; Aliasgharpour, H.; Mahmoudi, J.; Sadigh-Etemad, S.; Farhoudi, M.; Baniasadi, N. The efficacy of the ketogenic diet on motor functions in Parkinson’s disease: A rat model. Iran. J. Neurol. 2016, 15, 63–69. [Google Scholar]
  149. Zhao, Z.; Lange, D.J.; Voustianiouk, A.; MacGrogan, D.; Ho, L.; Suh, J.; Humala, N.; Thiyagarajan, M.; Wang, J.; Pasinetti, G.M. A ketogenic diet as a potential novel therapeutic intervention in amyotrophic lateral sclerosis. BMC Neurosci. 2006, 7, 29. [Google Scholar] [CrossRef]
  150. Mucke, L.; Selkoe, D.J. Neurotoxicity of amyloid β-protein: Synaptic and network dysfunction. Cold Spring Harb. Perspect. Med. 2012, 2, a006338. [Google Scholar] [CrossRef]
  151. Yoshikawa, K. Neurotoxicity of β-amyloid. Nature 1993, 361, 122–123. [Google Scholar] [CrossRef]
  152. Leng, F.; Edison, P. Neuroinflammation and microglial activation in Alzheimer disease: Where do we go from here? Nat. Rev. Neurol. 2021, 17, 157–172. [Google Scholar] [CrossRef] [PubMed]
  153. Heneka, M.T.; Carson, M.J.; El Khoury, J.; Landreth, G.E.; Brosseron, F.; Feinstein, D.L.; Jacobs, A.H.; Wyss-Coray, T.; Vitorica, J.; Ransohoff, R.M.; et al. Neuroinflammation in Alzheimer’s disease. Lancet Neurol. 2015, 14, 388–405. [Google Scholar] [CrossRef] [PubMed]
  154. Kim, D.Y.; Hao, J.; Liu, R.; Turner, G.; Shi, F.D.; Rho, J.M. Inflammation-mediated memory dysfunction and effects of a ketogenic diet in a murine model of multiple sclerosis. PLoS ONE 2012, 7, e35476. [Google Scholar] [CrossRef] [PubMed]
  155. Liu, H.; Huang, J.; Liu, H.; Li, F.; Peng, Q.; Liu, C. Effects of ketogenic diet containing medium-chain fatty acids on serum inflammatory factor and mTOR signaling pathway in rats. Chem. Biol. Technol. Agric. 2020, 7, 27. [Google Scholar] [CrossRef]
  156. McDaniel, S.S.; Rensing, N.R.; Thio, L.L.; Yamada, K.A.; Wong, M. The ketogenic diet inhibits the mammalian target of rapamycin (mTOR) pathway. Epilepsia 2011, 52, e7–e11. [Google Scholar] [CrossRef] [PubMed]
  157. Zhong, S.; Zhou, Z.; Lin, X.; Liu, F.; Liu, C.; Liu, Z.; Deng, W.; Zhang, X.; Chang, H.; Zhao, C. Ketogenic diet prevents paclitaxel-induced neuropathic nociception through activation of PPARγ signalling pathway and inhibition of neuroinflammation in rat dorsal root ganglion. Eur. J. Neurosci. 2021, 54, 5341–5356. [Google Scholar] [CrossRef] [PubMed]
  158. Simeone, T.A.; Matthews, S.A.; Samson, K.K.; Simeone, K.A. Regulation of brain PPARgamma2 contributes to ketogenic diet anti-seizure efficacy. Exp. Neurol. 2017, 287, 54–64. [Google Scholar] [CrossRef]
  159. Guo, M.; Wang, X.; Zhao, Y.; Yang, Q.; Ding, H.; Dong, Q.; Chen, X.; Cui, M. Ketogenic Diet Improves Brain Ischemic Tolerance and Inhibits NLRP3 Inflammasome Activation by Preventing Drp1-Mediated Mitochondrial Fission and Endoplasmic Reticulum Stress. Front. Mol. Neurosci. 2018, 11, 86. [Google Scholar] [CrossRef]
  160. Lin, J.; Huang, Z.; Liu, J.; Huang, Z.; Liu, Y.; Liu, Q.; Yang, Z.; Li, R.; Wu, X.; Shi, Z.; et al. Neuroprotective Effect of Ketone Metabolism on Inhibiting Inflammatory Response by Regulating Macrophage Polarization After Acute Cervical Spinal Cord Injury in Rats. Front. Neurosci. 2020, 14, 583611. [Google Scholar] [CrossRef]
  161. Haskó, G.; Cronstein, B. Regulation of inflammation by adenosine. Front. Immunol. 2013, 4, 85. [Google Scholar] [CrossRef] [PubMed]
  162. Koh, S.; Dupuis, N.; Auvin, S. Ketogenic diet and Neuroinflammation. Epilepsy Res. 2020, 167, 106454. [Google Scholar] [CrossRef] [PubMed]
  163. Yi, J.H.; Park, S.W.; Brooks, N.; Lang, B.T.; Vemuganti, R. PPARgamma agonist rosiglitazone is neuroprotective after traumatic brain injury via anti-inflammatory and anti-oxidative mechanisms. Brain Res. 2008, 1244, 164–172. [Google Scholar] [CrossRef] [PubMed]
  164. Croasdell, A.; Duffney, P.F.; Kim, N.; Lacy, S.H.; Sime, P.J.; Phipps, R.P. PPARγ and the Innate Immune System Mediate the Resolution of Inflammation. PPAR Res. 2015, 14, 549691. [Google Scholar]
  165. Bernardo, A.; Minghetti, L. PPAR-gamma agonists as regulators of microglial activation and brain inflammation. Curr. Pharm. Des. 2006, 12, 93–109. [Google Scholar] [CrossRef]
  166. Chawla, A. Control of macrophage activation and function by PPARs. Circ. Res. 2010, 106, 1559–1569. [Google Scholar] [CrossRef]
  167. Carvalho, M.V.; Gonçalves-de-Albuquerque, C.F.; Silva, A.R. PPAR Gamma: From Definition to Molecular Targets and Therapy of Lung Diseases. Int. J. Mol. Sci. 2021, 22, 805. [Google Scholar] [CrossRef]
  168. Mohamed, H.E.; El-Swefy, S.E.; Rashed, L.A.; Abd El-Latif, S.K. Biochemical effect of a ketogenic diet on the brains of obese adult rats. J. Clin. Neurosci. 2010, 17, 899–904. [Google Scholar] [CrossRef]
  169. Gzielo, K.; Soltys, Z.; Rajfur, Z.; Setkowicz, Z.K. The Impact of the Ketogenic Diet on Glial Cells Morphology. A Quantitative Morphological Analysis. Neuroscience 2019, 413, 239–251. [Google Scholar] [CrossRef]
  170. Attaye, I.; Van Oppenraaij, S.; Warmbrunn, M.V.; Nieuwdorp, M. The Role of the Gut Microbiota on the Beneficial Effects of Ketogenic Diets. Nutrients 2021, 14, 191. [Google Scholar] [CrossRef]
  171. Ang, Q.Y.; Alexander, M.; Newman, J.C.; Tian, Y.; Cai, J.; Upadhyay, V.; Turnbaugh, J.A.; Verdin, E.; Hall, K.D.; Leibel, R.L.; et al. Ketogenic Diets Alter the Gut Microbiome Resulting in Decreased Intestinal Th17 Cells. Cell 2020, 181, 1263–1275. [Google Scholar] [CrossRef] [PubMed]
  172. Ma, D.; Wang, A.C.; Parikh, I.; Green, S.J.; Hoffman, J.D.; Chlipala, G.; Murphy, M.P.; Sokola, B.S.; Bauer, B.; Hartz, A.M.S.; et al. Ketogenic diet enhances neurovascular function with altered gut microbiome in young healthy mice. Sci. Rep. 2018, 8, 6670. [Google Scholar] [CrossRef] [PubMed]
  173. Crosby, L.; Davis, B.; Joshi, S.; Jardine, M.; Paul, J.; Neola, M.; Barnard, N.D. Ketogenic Diets and Chronic Disease: Weighing the Benefits Against the Risks. Front. Nutr. 2021, 8, 702802. [Google Scholar] [CrossRef] [PubMed]
  174. Wang, D.; Mitchell, E.S. Cognition and Synaptic-Plasticity Related Changes in Aged Rats Supplemented with 8- and 10-Carbon Medium Chain Triglycerides. PLoS ONE 2016, 11, e0160159. [Google Scholar] [CrossRef] [PubMed]
  175. Pawlosky, R.J.; Kemper, M.F.; Kashiwaya, Y.; King, M.T.; Mattson, M.P.; Veech, R.L. Effects of a dietary ketone ester on hippocampal glycolytic and tricarboxylic acid cycle intermediates and amino acids in a 3xTgAD mouse model of Alzheimer’s disease. J. Neurochem. 2017, 141, 195–207. [Google Scholar] [CrossRef]
  176. Kashiwaya, Y.; Bergman, C.; Lee, J.H.; Wan, R.; King, M.T.; Mughal, M.R.; Okun, E.; Clarke, K.; Mattson, M.P.; Veech, R.L. A ketone ester diet exhibits anxiolytic and cognition-sparing properties, and lessens amyloid and tau pathologies in a mouse model of Alzheimer’s disease. Neurobiol. Aging 2013, 34, 1530–1539. [Google Scholar] [CrossRef]
  177. Kovács, Z.; D’Agostino, D.P.; Diamond, D.; Kindy, M.S.; Rogers, C.; Ari, C. Therapeutic Potential of Exogenous Ketone Supplement Induced Ketosis in the Treatment of Psychiatric Disorders: Review of Current Literature. Front. Psychiatry 2019, 10, 363. [Google Scholar] [CrossRef]
  178. Tabaie, E.A.; Reddy, A.J.; Brahmbhatt, H. A narrative review on the effects of a ketogenic diet on patients with Alzheimer’s disease. AIMS Public Health 2021, 9, 185–193. [Google Scholar] [CrossRef]
  179. Brandt, J.; Buchholz, A.; Henry-Barron, B.; Vizthum, D.; Avramopoulos, D.; Cervenka, M.C. Preliminary Report on the Feasibility and Efficacy of the Modified Atkins Diet for Treatment of Mild Cognitive Impairment and Early Alzheimer’s Disease. J. Alzheimers Dis. 2019, 68, 969–981. [Google Scholar] [CrossRef]
  180. Phillips, M.C.L.; Deprez, L.M.; Mortimer, G.M.N.; Murtagh, D.K.J.; McCoy, S.; Mylchreest, R.; Gilbertson, L.J.; Clark, K.M.; Simpson, P.V.; McManus, E.J.; et al. Randomized crossover trial of a modified ketogenic diet in Alzheimer’s disease. Alzheimers Res. Ther. 2021, 13, 51. [Google Scholar] [CrossRef]
  181. Neth, B.J.; Mintz, A.; Whitlow, C.; Jung, Y.; Solingapuram Sai, K.; Register, T.C.; Kellar, D.; Lockhart, S.N.; Hoscheidt, S.; Maldjian, J.; et al. Modified ketogenic diet is associated with improved cerebrospinal fluid biomarker profile, cerebral perfusion, and cerebral ketone body uptake in older adults at risk for Alzheimer’s disease: A pilot study. Neurobiol. Aging 2020, 86, 54–63. [Google Scholar] [CrossRef] [PubMed]
  182. Nagpal, R.; Neth, B.J.; Wang, S.; Craft, S.; Yadav, H. Modified Mediterranean-ketogenic diet modulates gut microbiome and short-chain fatty acids in association with Alzheimer’s disease markers in subjects with mild cognitive impairment. EBioMedicine 2019, 47, 529–542. [Google Scholar] [CrossRef] [PubMed]
  183. Taylor, M.K.; Sullivan, D.K.; Mahnken, J.D.; Burns, J.M.; Swerdlow, R.H. Feasibility and efficacy data from a ketogenic diet intervention in Alzheimer’s disease. Alzheimers Dement. 2017, 4, 28–36. [Google Scholar] [CrossRef] [PubMed]
  184. Ota, M.; Matsuo, J.; Ishida, I.; Takano, H.; Yokoi, Y.; Hori, H.; Yoshida, S.; Ashida, K.; Nakamura, K.; Takahashi, T.; et al. Effects of a medium-chain triglyceride-based ketogenic formula on cognitive function in patients with mild-to-moderate Alzheimer’s disease. Neurosci. Lett. 2019, 690, 232–236. [Google Scholar] [CrossRef] [PubMed]
  185. Rebello, C.J.; Keller, J.N.; Liu, A.G.; Johnson, W.D.; Greenway, F.L. Pilot feasibility and safety study examining the effect of medium chain triglyceride supplementation in subjects with mild cognitive impairment: A randomized controlled trial. BBA Clin. 2015, 3, 123–125. [Google Scholar] [CrossRef] [PubMed]
  186. Ohnuma, T.; Toda, A.; Kimoto, A.; Takebayashi, Y.; Higashiyama, R.; Tagata, Y.; Ito, M.; Ota, T.; Shibata, N.; Arai, H. Benefits of use, and tolerance of, medium-chain triglyceride medical food in the management of Japanese patients with Alzheimer’s disease: A prospective, open-label pilot study. Clin. Interv. Aging 2016, 11, 29–36. [Google Scholar] [CrossRef] [PubMed]
  187. Henderson, S.T.; Vogel, J.L.; Barr, L.J.; Garvin, F.; Jones, J.J.; Costantini, L.C. Study of the ketogenic agent AC-1202 in mild to moderate Alzheimer’s disease: A randomized, double-blind, placebo-controlled, multicenter trial. Nutr. Metab. 2009, 6, 31. [Google Scholar] [CrossRef] [PubMed]
  188. Torosyan, N.; Sethanandha, C.; Grill, J.D.; Dilley, M.L.; Lee, J.; Cummings, J.L.; Ossinalde, C.; Silverman, D.H. Changes in regional cerebral blood flow associated with a 45 day course of the ketogenic agent, caprylidene, in patients with mild to moderate Alzheimer’s disease: Results of a randomized, double-blinded, pilot study. Exp. Gerontol. 2018, 111, 118–121. [Google Scholar] [CrossRef]
  189. Reger, M.A.; Henderson, S.T.; Hale, C.; Cholerton, B.; Baker, L.D.; Watson, G.S.; Hyde, K.; Chapman, D.; Craft, S. Effects of beta-hydroxybutyrate on cognition in memory-impaired adults. Neurobiol. Aging 2004, 25, 311–314. [Google Scholar] [CrossRef]
  190. Fortier, M.; Castellano, C.A.; Croteau, E.; Langlois, F.; Bocti, C.; St-Pierre, V.; Vandenberghe, C.; Bernier, M.; Roy, M.; Descoteaux, M.; et al. A ketogenic drink improves brain energy and some measures of cognition in mild cognitive impairment. Alzheimers Dement. 2019, 5, 625–634. [Google Scholar] [CrossRef]
Figure 1. This schematic illustrates the synthesis and catabolism of ketone bodies within the body. (A) Ketones are produced in the liver under certain conditions. Consuming a high-fat low-carbohydrate diet promotes the production of ketone bodies as acetone, beta-hydroxybutyrate (BOHb) and acetoacetate in the liver. (B) Once in the bloodstream, free fatty acids from the ketogenic diet or adipocytes (lipolysis) enter the liver via the hepatic portal vein, where they participate in the ketogenesis process. Inside liver cells, free fatty acids access mitochondria via Carnitine Palmitoyltransferase (CPT) transporter channels. Within the mitochondria, fatty acids undergo oxidation, leading to the formation of essential ketone bodies. These ketone bodies exit the mitochondria through Monocarboxylate Transporter (MCT) channels, eventually leaving hepatocytes and entering the bloodstream, en route to the brain. (C) In the brain, ketone bodies enter the neuronal mitochondrial matrix through Monocarboxylate Transporters (MCT channels), where they undergo ketolysis. The action of Succinyl-CoA on ketone bodies promotes the formation of Acetyl-CoA, which subsequently integrates into the tricarboxylic acid (TCA) cycle for ATP generation. (D) Alzheimer’s disease is intricately linked to impaired glucose metabolism, resulting in an energy imbalance. Reduced glucose levels in cells compromise neuronal activity and metabolism. In this scenario, with diminished glucose availability, ketone bodies become the primary source for ATP production, allowing cells to maintain high energy levels and neuronal stability. This figure was partly generated using Servier Medical Art, provided by Servier, licensed under a Creative Commons Attribution 3.0 unported license.
Figure 1. This schematic illustrates the synthesis and catabolism of ketone bodies within the body. (A) Ketones are produced in the liver under certain conditions. Consuming a high-fat low-carbohydrate diet promotes the production of ketone bodies as acetone, beta-hydroxybutyrate (BOHb) and acetoacetate in the liver. (B) Once in the bloodstream, free fatty acids from the ketogenic diet or adipocytes (lipolysis) enter the liver via the hepatic portal vein, where they participate in the ketogenesis process. Inside liver cells, free fatty acids access mitochondria via Carnitine Palmitoyltransferase (CPT) transporter channels. Within the mitochondria, fatty acids undergo oxidation, leading to the formation of essential ketone bodies. These ketone bodies exit the mitochondria through Monocarboxylate Transporter (MCT) channels, eventually leaving hepatocytes and entering the bloodstream, en route to the brain. (C) In the brain, ketone bodies enter the neuronal mitochondrial matrix through Monocarboxylate Transporters (MCT channels), where they undergo ketolysis. The action of Succinyl-CoA on ketone bodies promotes the formation of Acetyl-CoA, which subsequently integrates into the tricarboxylic acid (TCA) cycle for ATP generation. (D) Alzheimer’s disease is intricately linked to impaired glucose metabolism, resulting in an energy imbalance. Reduced glucose levels in cells compromise neuronal activity and metabolism. In this scenario, with diminished glucose availability, ketone bodies become the primary source for ATP production, allowing cells to maintain high energy levels and neuronal stability. This figure was partly generated using Servier Medical Art, provided by Servier, licensed under a Creative Commons Attribution 3.0 unported license.
Metabolites 14 00025 g001
Table 1. Types of ketogenic diets.
Table 1. Types of ketogenic diets.
KD TypeRatioCarbohydrate Intake per Day on a Diet of 1000 KcalConsiderations
LCT4:1 to 3:18 g on a 4:1
16 g on a 3:1
Severe carbohydrate restriction, unpalatable
MCTDNot diet-ratio-related48 gMore ketogenic, gastrointestinal side effects
MADApproximately 1:140–60 gNo precise weighing, no protein/calorie restrictions
LOGIApproximately 1:110 g for the first month
then 20–30 g
Minimized glycemic increases, liberalized regimen
Abbreviations: classic long-chain triglyceride KD (LCT), medium-chain triglyceride KD (MCT), modified Atkins diet (MAD) and low glycemic index diet (LOGI).
Table 3. KD results in clinical trials of AD.
Table 3. KD results in clinical trials of AD.
AuthorParticipants (n)Primary DiagnosisDiet/
Supplementation
DurationKetosisResults/Side Effects
Croteau, E. et al. [93]15Mild-moderate ADMedium-chain triglyceride KD Suppl. C8C10 and C8Two periods of 1 monthC8C10: Blood BOHb (mM) = 0.46 ± 0.19
C8: Blood BOHb (mM) = 0.57 ± 0.27
Increased total brain energy metabolism.
Phillips, M. C. L. et al. [180]26Alzheimer’s disease Classic KDTwo periods of 12 weeksBOHb level = 0.95 ± 0.34 mmol/L
18/21 patients
Improved daily function and quality of life.
No significant changes were observed in the lipid profile.
Mild adverse effects.
Neth, B. J. et al. [181]20Subjective memory complaints or mild cognitive impairmentMediterranean KDTwo periods of 6 weeksBOHb level = 0.23 (0.27) mmol/LImprovement of peripheral metabolic measures, CSF biomarker profile and increased cerebral perfusion.
No serious adverse events occurred.
Nagpal, R. et al. [182]17Mild cognitive impairment or cognitively normalMediterranean KDTwo periods of 6 weeksNot measuredModulating capacity of MMKD in the gut microbiome.
Taylor, M. K. et al. [183]15Very-mild, mild and moderate ADMedium-chain triglyceride KD3 monthsSerum BOHb level = 0.31 mmol/LOverall cognitive improvement.
No serious adverse events occurred.
Ota, M. et al. [184]20Mild-moderate ADMedium-chain triglyceride KD12 monthsPlasma BOHb level = 470.9 ± 292.6 μmol/LPositive effects on verbal memory and processing speed.
Diarrhea, most frequently reported side effect of MCT.
Rebello, C. J. et al. [185]6Mild cognitive impairmentMedium-chain triglyceride KD24 weeksApoE4(−) Serum BOHb level = 0.15 mM
ApoE4(+) Serum BOHb level = 0.54 mM
Memory improvement.
Ohnuma, T. et al. [186]22Sporadic, mild-moderate ADAXONA Dietary Suplment3 monthsSerum BOHb level = 81.1 ± 79.9 μMImprovement in cognitive functions.
No severe gastrointestinal adverse effects.
Henderson, S. T. et al. [187]152Mild-moderate ADAC-120290 daysBOHb level = 0.39 mMCognitive improvement.
Mild to moderate adverse events restricted to the gastrointestinal system.
Torosyan, N. et al. [188]16Mild-moderate ADCaprylidene45 daysNot measuredIncreased blood flow in specific brain regions.
Reger, M. A. et al. [189]20AD or mild cognitiveEmulsified medium-chain triglycerideTwo times of 90 minApoE4(−) Serum BOHb level = 0.54 mM
ApoE4(+) Serum BOHb level = 0.43 mM
Cognitive improvement in AD patients withoutAPoE4.
Fortier, M. et al. [190]52Mild cognitive impairmentKetogenic medium-chain triglyceride drink6 monthsPlasma BOHb level = 401 ± 303 μMImproved several cognitive outcomes in MCI.
No severe adverse events.
Abbreviations: ApoE4 = apolipoprotein E4; BOHb = beta-hydroxybutyrate; CSF = cerebral spinal fluid. +: plus APOE4; −: without APOE4.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Oliveira, T.P.D.; Morais, A.L.B.; dos Reis, P.L.B.; Palotás, A.; Vieira, L.B. A Potential Role for the Ketogenic Diet in Alzheimer’s Disease Treatment: Exploring Pre-Clinical and Clinical Evidence. Metabolites 2024, 14, 25. https://doi.org/10.3390/metabo14010025

AMA Style

Oliveira TPD, Morais ALB, dos Reis PLB, Palotás A, Vieira LB. A Potential Role for the Ketogenic Diet in Alzheimer’s Disease Treatment: Exploring Pre-Clinical and Clinical Evidence. Metabolites. 2024; 14(1):25. https://doi.org/10.3390/metabo14010025

Chicago/Turabian Style

Oliveira, Tadeu P. D., Ana L. B. Morais, Pedro L. B. dos Reis, András Palotás, and Luciene B. Vieira. 2024. "A Potential Role for the Ketogenic Diet in Alzheimer’s Disease Treatment: Exploring Pre-Clinical and Clinical Evidence" Metabolites 14, no. 1: 25. https://doi.org/10.3390/metabo14010025

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop