Next Article in Journal
Echinacea purpurea Fractions Represent Promising Plant-Based Anti-Inflammatory Formulations
Next Article in Special Issue
The Chemical Variability, Nutraceutical Value, and Food-Industry and Cosmetic Applications of Citrus Plants: A Critical Review
Previous Article in Journal
The Impact of Za’atar Antioxidant Compounds on the Gut Microbiota and Gastrointestinal Disorders: Insights for Future Clinical Applications
Previous Article in Special Issue
Characterization and Preliminary In Vitro Antioxidant Activity of a New Multidrug Formulation Based on the Co-Encapsulation of Rutin and the α-Acylamino-β-Lactone NAAA Inhibitor URB894 within PLGA Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Honey and Alzheimer’s Disease—Current Understanding and Future Prospects

1
Department of Anatomy, Faculty of Medicine, Universiti Kebangsaan Malaysia, Kuala Lumpur 56000, Malaysia
2
Department of Physiology, Faculty of Medicine, Universiti Kebangsaan Malaysia, Kuala Lumpur 56000, Malaysia
*
Author to whom correspondence should be addressed.
Antioxidants 2023, 12(2), 427; https://doi.org/10.3390/antiox12020427
Submission received: 30 December 2022 / Revised: 30 January 2023 / Accepted: 3 February 2023 / Published: 9 February 2023
(This article belongs to the Special Issue The 10th Anniversary of Antioxidants: Past, Present and Future)

Abstract

:
Alzheimer’s disease (AD), a leading cause of dementia, has been a global concern. AD is associated with the involvement of the central nervous system that causes the characteristic impaired memory, cognitive deficits, and behavioral abnormalities. These abnormalities caused by AD is known to be attributed by extracellular aggregates of amyloid beta plaques and intracellular neurofibrillary tangles. Additionally, genetic factors such as abnormality in the expression of APOE, APP, BACE1, PSEN-1, and PSEN-2 play a role in the disease. As the current treatment aims to treat the symptoms and to slow the disease progression, there has been a continuous search for new nutraceutical agent or medicine to help prevent and cure AD pathology. In this quest, honey has emerged as a powerful nootropic agent. Numerous studies have demonstrated that the high flavonoids and phenolic acids content in honey exerts its antioxidant, anti-inflammatory, and neuroprotective properties. This review summarizes the effect of main flavonoid compounds found in honey on the physiological functioning of the central nervous system, and the effect of honey intake on memory and cognition in various animal model. This review provides a new insight on the potential of honey to prevent AD pathology, as well as to ameliorate the damage in the developed AD.

1. Introduction

Alzheimer’s disease (AD) is a neurodegenerative disorder associated with damage to the brain areas such as the cerebral cortex, temporal lobe, hippocampus, amygdala, entorhinal cortex (EC), and parahippocampal region [1,2]. The disease is mainly characterized by impaired memory and cognitive deficits [3,4]. AD is considered the most common cause of dementia, accounting for about 60–70% of the total cases worldwide [5]. In addition to the deficits of memory and cognition, AD is also accompanied by behavioral changes. Since most of the areas affected by the pathology are involved both in cognition and behavior, the predominant behavioral changes, such as agitation, dysphoria, and apathy, are correlated highly with cognitive dysfunction [6].
Previous studies have proposed several effective solutions to reduce the deposition of amyloid fibrils, minimize oxidative stress and neuroinflammation, and/or improve memory and cognition. These can be divided into drugs and antioxidants or neuroprotective agents for ease of discussion. The drugs include N-methyl-d-aspartate (NMDA) receptors antagonists [7,8], agents acting on the acetylcholinergic system (ACh system) [9,10], anti-amyloid [11], and anti-tau [12]. The antioxidants or neuroprotective agents include idebenone (an organic compound from the quinone family) and α-tocopherol [13], estrogen analogues [14,15], and honey [16,17,18]. Although the allopathic medications have shown promising results in attenuating symptoms, they may cause a number of adverse effects while having some serious precautions and contraindications [10,19,20,21,22,23,24,25]. Comparably, honey only has one serious adverse effect that ranges from mild hypersensitivity reaction to anaphylactic shock, which is attributed to the presence of pollens and bee-derived proteins in honey [26]. These allergic reactions, however, are very rare with only a small number of cases reported till date [26,27,28]. Additionally, among all the above mentioned substances, honey stands out by having the potential to improve almost all aspects of AD, such as oxidative stress [29,30,31], neuroinflammation [32,33], neuroprotection [34,35], ACh system [36], and memory and cognition [37,38].
As AD is associated with the involvement of the central nervous system (CNS) that causes the characteristic impaired memory, cognitive deficits, and behavioral abnormalities, in this review article, we will limit our discussion of the effects of honey on the said aspects.

2. Pathophysiology and Clinical Picture of Alzheimer’s Disease

AD is believed to begin and caused by the accumulation of amyloid beta (Aβ) plaques; this perspective of AD progression is known as the amyloid-cascade-hypothesis [39]. According to this hypothesis, the neuropathology in AD starts from the extracellular accumulation of Aβ fibrils as abnormal neuritic plaques, the deposition of which leads to oxidative damage and inflammation. Although this is a widely accepted hypothesis, some researchers believe in the tau hypothesis, according to which tau pathology is a prerequisite for the Aβ aggregation to take place [40,41,42]. Additionally, there is a third viewpoint which suggests that there may be more than one pathological pathway co-occurring, as dementia in AD is not correlated with either plaque or tangle burden but with the serum amyloid protein content in the Aβ plaques [43,44]. This hypothesis is further supported by the findings that the neurofibrillary tangles (NFT)-bearing neocortical neurons are functionally intact [45], and even though the cognitive deficits increase with ageing, the load of neuritic plaques and NFTs tend to decline as the elderly people age, i.e., more burden in the 60–80-year-old individuals than in over 90-years old individuals [46].
Pathologically, AD is characterized by the deposition of Aβ plaques (extracellularly) and NFTs (intracellularly). Soon after the tau fibrils are hyperphosphorylated, they may be converted into pathological tau and result in the formation of NFTs [47]; the latter more commonly affects the medial limbic structures (MLS) comprised of hippocampus, subiculum, EC, and amygdala [48,49]. The tau aggregates need the presence of neuritic plaques, therefore, are likely formed adjacent to them [47], whereas NFTs have an independent presence [49]. Irrespective of the site of impaction of tangles, which may vary in the brain, the formation of the NFTs is possibly the result of the interplay of oxidative injury, neuroinflammation, ineffective degradation, and subsequent ubiquitination causing hyperphosphorylation of tau followed by the subsequent formation of tangles [50,51,52,53].
Considering the defect at the genetic level, AD results from the abnormality in the expression of five genes: Apolipoprotein E (APOE), Amyloid Precursor Protein (APP), Beta-site Amyloid precursor protein Cleaving Enzyme 1 (BACE1), Presenilin 1 and 2 (PSEN-1, and PSEN-2). While the pathology of the first gene APOE (especially the allelic variant ε4) is associated with sporadic AD [54], the latter four genes were found to be responsible for the familial AD. The most common form of AD is sporadic and its risk increases with the presence of ε4 allele [55,56]. The other alleles, i.e., ε2 and ε3, minimize oxidative damage and neuronal death, whereas, the ε4 allele has the lowest capacity to prevent cellular toxicity [57]. Therefore, its presence increases the likelihood of developing AD [57,58]. As for the familial AD, the APP gene is located on chromosome 21 and is responsible for the production of APP, which is required for the normal regulation of several cellular functions [59]; however, excessive dose, hence over-expression, of this gene results in increased amyloid levels in brain and likelihood to develop AD [60,61], as also observed in Down syndrome (trisomy 21) [62]. The other genes, BACE1 and PSEN (1 and 2), also known as β-secretase and γ-secretase, further play their part in AD pathogenesis [63,64,65]. It occurs when the APP is cleaved by BACE1 (β- secretase) instead of the normal cleavage by α-secretase, and the product acts as a substrate for γ-secretase resulting in the formation, and subsequently, aggregation of the Aβ oligomers [66,67]. Furthermore, since the β-secretase and γ-secretase act on the common substrate, the elevated level and activity of the former is mostly accompanied by a reduction in the level of the latter [68].
Overall, the damage in the brain in AD is comprised of injuries on both macroscopic and microscopic levels. The gross changes consist of a reduction of the total brain tissue with an increase in the volume of the ventricles [69,70], whereas the underlying microscopic changes include the loss of synapses [71,72], damage to pyramidal neurons, and neurodegeneration [1,73]. In addition, the loss of synapses can either occur in the presence of normal long-term potentiation (LTP) [74] or is probably due to impaired LTP [75,76]. To further shed the light on these contrasting results, recent studies described that AD may affect LTP in some pathways (e.g., Schaffer collateral) while the LTP in other pathways (e.g., mossy fibers) remain unaffected/normal [77] with a possible alteration in the short-term potentiation [78]. Moreover, AD brains are affected by oxidative injury and inflammatory damage. The former is due to an imbalance between antioxidants and oxidation-causing substances (i.e., free radicals and reactive oxygen species), causing a reduction in the activity of antioxidants such as superoxide dismutase (SOD), glutathione (GSH), and catalase, together with an increase in the markers of oxidative damage such as Malondialdehyde (MDA) (the product of lipid peroxidation) and 3-nitrotyrosine (the end product of protein oxidation), and 8-hydroxydeoxyguanosine and 8-hydroxyguanosine (the product of oxidation of guanine in DNA) [79,80,81]. Likewise, perpetual neuroinflammation marked by an imbalance in the inflammatory cytokines is also evident by the over-expression of the pro-inflammatory markers, such as IL-1α, IL-1β, IL-6 [82,83], TNFα, and NFκB [84,85], and an accompanied under-expression of some anti-inflammatory cytokines, such as IL-4 and IL-10 [86,87]. Moreover, the reduced level of anti-inflammatory cytokines further leads to the uninhibited activity of pro-inflammatory cytokines [87,88,89] and results in more neuronal damage [87]. Furthermore, an altered interaction of cytokines (both pro-inflammatory and anti-inflammatory) has been observed based on the underlying pathology of APOE genotypes [90,91].
Those mentioned structural and functional abnormalities that are the characteristic features of AD can start appearing in the brain in middle-aged individuals (familial or early-onset AD) or the elderly (sporadic or late-onset AD). Irrespective of the age of onset, clinically, AD presents as deficits of memory, cognition [48,92], and behavior [6].

3. Honey and Its Powerful Ingredients—The Phenolic Compounds

Honey mainly contains sugar and water [93]. The high sugar content, comprised of dextrose, levulose, and other complex carbohydrates, makes it a better alternative to glucose as it replenishes energy with a constant blood glucose level [94]. In addition to being a mixture of around 30 different kinds of sugars [95], honey has several minor components, including phenolic compounds, proteins, amino acids, vitamins, enzymes, and minerals [93,96]. Although the main constituents (water and sugars) remain the same, the composition of minor components of each honey type varies significantly, which is due to the difference in geographical location, floral source, storage, and the final color [95,97]. Due to the mentioned factors, various types of honey are different in composition of polyphenolic compounds [98,99], and therefore, polyphenolic activity and total antioxidant capacity (TAC). The quantification of total phenolic content showed that certain types of honey, such as, stingless bee honey and Tualang honey have higher content of phenolic acids and flavonoids, and greater TAC and radical scavenging activity [100,101,102,103] which may indicate more potential in attenuation of oxidative stress in vivo as well. To our knowledge, presently, no study has been conducted comparing antioxidant effects of various types of honey in vivo. Moreover, the composition of the same variant of honey has not been compared from different regions around the world so far, which points toward a likelihood of varied composition of honey obtained from two different countries.
Studies suggest that most of honey’s antioxidant, anti-inflammatory, and neuroprotective properties are due to its phenolic content [104,105]. Phenolic compounds are comprised of four classes of polyphenols: Phenolic acids, flavonoids, stilbenes, and lignans. Out of these, phenolic acids and flavonoids primarily have the potential to act as antioxidants and reduce oxidative stress [29,30,31] and neuroinflammation [32] that are the mediators of insults to the brain in the neurodegenerative diseases [106,107]. However, this review will focus on the effectiveness of flavonoids and phenolic acids on the CNS and in the prevention/treatment of AD pathology. The main phenolic compounds affecting the physiological functioning and/or the pathophysiology of the CNS are stated in Figure 1.
According to the studies on AD models (Refer to Table 1 and Table 2), all mentioned flavonoids and phenolic acids exert antioxidant effects and show neuroprotective activity. All agents, except myricetin, were also found to exhibit anti-inflammatory potential. As myricetin possesses an anti-inflammatory ability against post-ischemic neurodegeneration [108], if tested, it may also display similar potential in AD brain. In addition to the antioxidant and anti-inflammatory potential, most polyphenolic components also proved to attenuate AD pathology by decreasing amyloid deposition, with an exception of kaempferol and chlorogenic acid. Additionally, naringenin, naringin, quercetin, caffeic acid, and ellagic acid also reduce levels of p-tau in AD brain.

4. Therapeutic Potential of Flavonoids and Phenolic Acids

Several polyphenols are known to exert protective effects on the nervous system and are suggested to have a role in alleviating symptoms of neurological diseases [159,160,161] including AD [162]. All these phenolic compounds, which are listed in Figure 1, are found to improve cognitive performance in AD pathology, and prevents from cognitive decline when ingested before inception of disease. These compounds are found in different kinds of honey in addition to other sources, and are particularly effective in attenuating oxidative stress along with exerting preventive effects on several other mechanisms in AD pathology. Their potential to reduce AD-induced brain injury is further proven by the microscopic studies and brain assays where they showed neuroprotective effects on the cortex [116,140], hippocampus [124,125], and hypothalamus [119]. Moreover, the consumption of polyphenols resulted in the prevention of hypoperfusion injury [145] with a generalized increment in cell number having normal physiology in the subiculum [122], and hippocampal proper area CA1 [48,136,139,154], CA3, and dentate gyrus [109,120,157], along with preserving normal synapses [112,131,144], the latter is further evident by an increased LTP after polyphenol consumption [154,156,158,163]. Further, the detailed studies of AD-brain animal model showed the potential of the polyphenols to reduce oxidation markers such as MDA [128,153], nitric oxide (NO), and nitrite [112,119,129,130], thereby attenuating free-radical-induced oxidation insults. The observed levels of antioxidants, however, are contrasting, with many studies deducing an elevated level of SOD and catalase, GSH [113,127,140,143,152,155], whereas others concluding decreased expression [115,146,148]. Although the results are contradictory, the studies demonstrating a reduction in the activity of antioxidants claim that this decrease also signifies the attenuation of oxidative injury, which subsequently renders the expression of the anti-inflammatory markers unnecessary. However, despite the discrepancy in the results, all studies conclude that the changes in the expression of these markers lead to reduced oxidative damage and Aβ-plaque accumulation.
Moreover, phenolic compounds can also alter the expression of some critical genes: APP, BACE1, PSEN-1, and Glutathione peroxidase 1 (GPx1). Polyphenols are found to down-regulate the expression of APP [164] and PSEN-1 gene [165], along with either a decrease [142,146,148] or increase in GPx1 expression [116]. As GPx1 is an enzyme that catalyzes the reduction of hydroperoxides and hydrogen peroxide by GSH to attenuate oxidation [166], reduced expression can lead to oxidative injury to cell. Moreover, the expression of BACE1 is also decreased [153] and is thought to be inhibited post-transcription, probably at the protein level [148]. Although not studied yet, a similar down-regulation of PSEN-2 gene expression can be expected by polyphenol consumption [167]. Moreover, polyphenols may also prevent neuritic plaque deposition by increasing α-secretase activity and by reducing cleavage of the APP to amyloidogenic soluble APP-β and β-CTFs [142] and hence, preventing the accumulation of the latter in synapses [148,165]. Taken together, these studies suggest that the polyphenols likely regulate gene expressions to reduce oxidation and formation of Aβ fibrils. Moreover, the polyphenols also increase the expression of the transcription factor Nrf2, which is responsible for regulating the induction of antioxidant genes, thereby improving defense against oxidative injury [139]. To protect the CNS further, the polyphenols reduce the level of pro-inflammatory markers, such as NFκB, TLR4 [139,157], COX-2 [151], MHC class II, TNFα [114,127,146], IL-1α [149], IL-1β [147,151,168], IL-6 [157,169] and increase anti-inflammatory cytokines [115], thereby reducing neuroinflammation. Furthermore, by decreasing neuroinflammation, these substances also attenuate the immunoreactivity of microglia and astroglia in the hippocampus, EC, and amygdala, which is commonly observed in AD neuropathology [115,122,142,146,149,150].
Additionally, the polyphenols reduce tau hyperphosphorylation and subsequent formation of NFT [170] and decrease the deposition of Aβ-plaques [140,156,171]. They also seem to exert a neuroprotective effect by preventing neuronal injury [136,152] and apoptosis [119,136,151] and by regulating the ACh system, where they increase ACh and choline acetyltransferase (ChAT), and decrease acetylcholinesterase (AChE) [119,123] and butyrylcholinesterase (BChE) [153]. These polyphenols’ effects also lead to minimization of deficits of memory and cognitive [172,173,174,175]. As honey contains a number of these polyphenols, its consumption can be expected to have similar potential to prevent and treat CNS pathology in AD. The therapeutic potential of the polyphenols: flavonoids and phenolic acids, are summarized in Table 1 and Table 2, respectively.
The effectiveness of honey in minimizing neurodegeneration is attributed to its neuroprotective effects on the brain [30,34], including the prefrontal cortex [176,177,178] and hippocampus [34,35,179,180]. Honey prevents neurodegeneration by attenuating two main phenomena, which are oxidative stress and neuroinflammation [36,132]. The reduction in neuroinflammation [181] is due to the attenuation of oxidative stress [30,31] and the prevention of free radical-mediated injury to the brain tissue [182,183]. This effect is evident by an increase in the antioxidant enzyme such as SOD and a reduction in the oxidative-stress markers, such as plasma MDA and protein carbonyl in aged brains [176,177]. Subsequently, as the hippocampal pyramidal neurons are highly susceptible to oxidative damage, this reduction of oxidative stress probably rescues them from insult and degeneration [34,182].
Further, along with the hippocampus, injury to the medial prefrontal cortex (mPFC) and piriform cortex is commonly observed in AD, both of which are associated with memory and cognition. Unlike other primary sensory cortices, which are minimally affected by the AD pathology, the piriform cortex is possibly affected even before or along with the development of the cognitive symptoms [184,185,186]. Hence, it is also considered a predictive marker of the conversion to AD [184,187]. Similarly, the damage to the mPFC in AD is evident as defective functioning [188,189] and abnormal connectivity with other associated brain areas [190]. Even though no such study has been undertaken to look at the injuries in these areas, the neuroprotective effects of honey may also rescue the mPFC and piriform cortical injury.
Although researchers widely accept the neuroprotective capacity of honey, we still do not have much data on the effects of honey on the physiology and/or anatomy of the human brain, and, therefore, its potential to act on the CNS is not fully understood to date. It is probably due to the late advent of technologies to study the brain in honey-related research and the limitation to researching human CNS. However, to overcome the limitation of experimental access to the human brains and to understand the possible effect of honey on the microscopic level, the research is now predominantly being carried out in rodents.

5. Effects of Honey on Memory, Cognition, and Behavior

Cortical Aβ deposition exerts effects on temporal lobe atrophy and resultant cognitive impairment in individuals with AD [191]. From psychophysiological perspective, cognition, learning, and memory are believed to be mainly determined by the cortico-hippocampal (C-H) circuit’s normal functioning [192,193]. As ACh is the principal neurotransmitter in synapses, the amount of ACh also plays an essential role in learning behavior and cognitive performance [194]. Although relatively constant, the amount of ACh still normally fluctuates, according to the need in the memory processes, such as encoding and retrieval [195]. Essentially, since the integrity of the C-H circuit depends upon the normal physiology of neurons and synapses, the Aβ plaque formation, and therefore AD, may affect the circuit by damaging neurons [196], reducing the number of cholinergic neurons [197], decreasing the ChAT activity [198,199], decreasing ACh release [200], and impairing synapses, which results in defective transmission [72,74]. Surprisingly, ageing and Aβ fibrillogenesis also decrease the AChE activity [201,202]; this finding is unexpected and in contrast to the decreased ACh indicates that the reduction in ACh is likely due to degeneration of the cholinergic neurons along with an increase in another cholinesterase enzyme activity, such as BChE [203] and not due to elevated AChE levels, as the latter is itself hydrolyzed by the former [204,205]. The intake of honey is found to reduce the level of BChE with a further decline in the level of AChE [176,206]. Although the exact mode of action is still not understood, this cholinesterase inhibition, together with neuroprotection, results in improved cognition and memory [207] after honey consumption, as observed in rodents [180,182,208,209] and humans [37,210]. The effects of honey as a nutraceutical agent in improving memory and cognition are further discussed in Table 3 (in rodents) and Table 4 (in humans).

6. Honey on Dopaminergic Neurons—Important Players in Memory Deficits in AD

In addition to ACh, dopamine plays a significant role in learning and memory functions. Besides being secreted from dopaminergic neurons (DN) of the Ventral tegmental area (VTA) and substantia nigra pars compacta (SNpC) in the midbrain [216,217], dopamine is also released by locus coeruleus (LC), located in the brainstem, which co-releases dopamine along with noradrenaline [218,219]; this released dopamine from LC innervates CA3 [220], and is thought to be the primary source of supply to the dorsal hippocampus [218,221]; however, a recent study suggests that the midbrain modulate dopaminergic innervation to the dorsal hippocampus and this stimulation is sufficient to arouse aversive memory even in the absence of input from LC [216]. To aid with the understanding of the contrasting source of dopamine, Takeuchi et al. proposed that although the projection of dopaminergic fibers from LC is denser than the midbrain [221], the midbrain and LC both modulate dorsal hippocampus in different kinds of memory consolidation processes [222]. Since the DN in the VTA are the primary site for dopamine synthesis, the DN in the midbrain-hippocampal (M-H) loop are vital in learning, memory formation, and consolidation [223,224]. The DN, and the secreted dopamine, modulate synaptic plasticity and contribute to the LTP in the hippocampus [225], thereby playing an essential part in the genesis and fortification of the episodic [226], aversive [216], and spatial memories [224]. Moreover, dopamine, together with norepinephrine, is crucial for the recognition memory [227,228]. In non-diseased brains, the number of DN and, therefore, the functional connectivity of the midbrain tends to decline with age [229,230], which may appear as deficits in learning and memory [231]. Similarly, as AD is a disease of old age, there is degeneration of DN [232,233]; However, due to the Aβ pathology, probably more damage occurs to the dopaminergic synapses in the M-H loop. Due to the mentioned insult, there is a decrease in dopamine, leading to impaired synaptic plasticity [234,235] and deficits of memory [233,236]. Polyphenols are found to prevent the degeneration of dopaminergic neurons and increase dopamine levels [137,237,238,239,240,241]. Although all these studies discuss the attenuation of neuroinflammation with/without Parkinson’s disease, similar results are expected in the AD model. As the insults to DN and reduced dopamine in AD are recently being studied in detail, more research is encouraged to be conducted on the AD M-H loop to understand the effect of honey and its constituent polyphenols on memory improvement in AD.

7. Honey as a Nootropic Agent—Prevention, Treatment, or Both?

In light of the previous research, it is evident that by acting on the CNS and working through various mechanisms, honey acts as a nootropic agent (refer Figure 2 and Figure 3). Now the question arises of the right time to utilize these nootropic properties to alleviate AD symptoms. Another similar issue is understanding whether honey consumption is effective in preventing the development/conversion of mild cognitive impairment into AD, mitigating the damage during ongoing AD disease pathology or reversing the injury done to the brain by AD. Although, to our knowledge, this aspect is not assessed to date, the studies on polyphenols (discussed in Table 1 and Table 2) may suggest some possible effects. The intake of phenolic compounds before initiation of the AD neuropathology is found to halt the progression of the CNS disease, protect neurons, reduce neuroinflammation and oxidative damage, and minimize memory and cognitive deficits, as seen in the studies on the AD-rodent models [48,109,111,112,117]. Likewise, honey ingestion in subjects with developed AD may also cause effects similar to those observed in the polyphenols-treated AD model [121,122,125,126]. Since these polyphenols are abundant in honey, we can expect the same benefits with honey consumption in human subjects.
Although honey is loaded with various kinds of polyphenols [100,102,103], the protective or curative effect of honey can be enhanced further by consuming it in combination with some nutraceutical agent [182,210,242]. Moreover, the synthesis of dimer by combining caffeic acid and ferulic acid [243], and the use of an amino acid (glutamine) conjugated with phenolic acid [244], both of which proved to be more efficient than the polyphenol alone, have paved a path for the likelihood of the advent of similar new combinations with honey that might emerge as the novel therapies for the prevention of AD. Moreover, a mixture of honey with other nutraceutical substances has proven to be effective in AD (for review: [245,246,247]) and can be appraised in prevention and/or management of AD.
In the same notion, the accurate dose of honey to prevent and/or treat AD has not been deduced to date. One of the important reasons of inability to draw conclusions is the fact that most studies are conducted on rodents, with very few studies on human subjects. Moreover, many confounding factors need to be addressed, such as the type of honey to be ingested, the therapeutic dose of honey, the minimum duration of honey intake, stage of AD (if given for treatment). Since few studies mentioned in Table 4 use a formulation of honey and other nootropic agents, another question arises whether combinations with such agents are more effective in terms of dosage and duration in improving human cognition. To our knowledge, currently there are no known studies on primates that observe the benefits of honey on cognition, or sporadic AD which is best modeled by the rhesus monkeys [248]. Comparative studies are encouraged in these areas, with possible usage of primates such as chimpanzees etc., to observe and deduce invaluable conclusions.

8. Conclusions and Future Directions

The phenolic compounds prevent damage to the neurons while promoting apoptosis in dysfunctional or cancer cells, which points toward different mechanisms of action in the brain cells than the rest of the body. On the same notion, the polyphenols are believed to exhibit oxidation-promoting properties for review: [249], rendering it necessary to explore the reasons for the switch between pro-oxidant and antioxidant characteristics to utilize these novel qualities appropriately. However, as various polyphenols have anti-amyloid and anti-tau potential, there is a possibility that they function as antioxidants in the cells that contain (or may contain) the Aβ aggregates and NFTs, whereas they promote oxidation in other abnormal cells; this aspect of these substances, as well as that of honey, needs further elaboration.
Some recent studies have demonstrated the potential of flavonoids in prevention of memory decline in elderly individuals [250,251,252,253]. Similarly, the research on animal models of AD (mentioned in Table 1 and Table 2) have shown an effectiveness of polyphenols in prevent and/or treat AD symptoms. However, the source of phenolic compounds in all these studies are variable. Considering the fact that the phenolic compounds can be obtained from various sources, notably fruits, vegetables, beverages, and nuts [254,255], the same polyphenols from distinct sources may have different pharmacological activities, and their polyphenolic potential, especially the capacity to act as an anti-oxidant and anti-inflammatory agent, may also vary accordingly. In the light of these considerations, new studies focusing primarily on the flavonoids and phenolic acids derived from honey are highly encouraged to understand the polyphenolic potential of honey.
Although an increased level of AChE is observed in cerebrospinal fluid (CSF) of AD patients [256], an overall reduction of AChE concentration is found in the AD brain [201,202]. The reduced amount of AChE suggests that AD pathology can be due to the reduction of some neuroprotective variant of AChE (e.g., AChE-R) in the absence of an increase in AChE activity [202]. Taking this into consideration, although the most reliable drug to treat mild to moderate AD is donepezil, an AChE inhibitor, it paradoxically increases the amount of AChE in the CSF [257,258]. These findings suggest that the AChE could have different properties inside the brain and within the CSF, pointing toward the possibility that the effect of honey on memory and cognition is due to neuroprotection with/without some mechanism other than AChE inhibition.
The role of dopaminergic system has been studied for a long time; however, its importance in memory deficits in AD was not clear. With the new studies on the role of dopaminergic system in learning and memory, a decline in dopamine levels with the damage in synapses is observed in AD. Furthermore, it is found that restoration of the dopaminergic levels is associated with improvement of memory deficits in AD [259]. For this purpose, dopamine agonists are being tested and proven to reverse the memory-related symptoms [260,261]. Moreover, the dopamine and its derivatives are found effective to reverse oxidative stress, inflammation, and Aβ load [262]. However, the dopamine replacement methods have given promising results, the effects of polyphenols and honey are still not being elucidated. Considering the polyphenol composition of honey, it can be more effective in treating various aspects of AD neuropathology compared to dopamine alone. On the same note, despite the fact that honey consumption is found to have a miraculous role in treating various diseases [263,264,265], its effectiveness in neurodegenerative diseases is still under evaluation. Having being proven to positively affect cognition and memory, the antioxidant capacity of honey also suggests its potential to manage neurological disorders and neurodegenerative diseases. More studies are needed to be conducted using animal models of neurodegenerative diseases, such as AD, to study the benefits of honey in its treatment and management.

Author Contributions

Conceptualization, A.S. and M.F.Y.; writing—original draft preparation, A.S.; writing—review and editing, M.F.Y., F.A., J.K., and S.L.T.; supervision, M.F.Y., F.A., J.K. and S.L.T.; project administration, M.F.Y.; funding acquisition, M.F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by UKM Research University Grant, grant number GUP-2021-038.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. da Silva Filho, S.R.B.; Barbosa, J.H.O.; Rondinoni, C.; dos Santos, A.C.; Salmon, C.E.G.; da Costa Lima, N.K.; Ferriolli, E.; Moriguti, J.C. Neuro-Degeneration Profile of Alzheimer’s Patients: A Brain Morphometry Study. NeuroImage Clin. 2017, 15, 15–24. [Google Scholar] [CrossRef] [PubMed]
  2. Klein-Koerkamp, Y.; Rolf, A.H.; Kylee, T.R.; Moreaud, O.; Keignart, S.; Krainik, A.; Hammers, A.; Baciu, M.; Hot, P.; Alzheimer’s Disease Neuroimaging Initiative. Amygdalar Atrophy in Early Alzheimer’s Disease. Curr. Alzheimer Res. 2014, 11, 239–252. [Google Scholar] [CrossRef] [PubMed]
  3. DeTure, M.A.; Dickson, D.W. The Neuropathological Diagnosis of Alzheimer’s Disease. Mol. Neurodegener. 2019, 14, 32. [Google Scholar] [CrossRef] [PubMed]
  4. Kumar, A.; Sidhu, J.; Goyal, A.; Tsao, J.W. Alzheimer Disease. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  5. WHO. Dementia. Available online: https://www.who.int/news-room/fact-sheets/detail/dementia (accessed on 16 November 2022).
  6. Mega, M.S.; Cummings, J.L.; Fiorello, T.; Gornbein, J. The Spectrum of Behavioral Changes in Alzheimer’s Disease. Neurology 1996, 46, 130–135. [Google Scholar] [CrossRef]
  7. Reisberg, B.; Doody, R.; Stöffler, A.; Schmitt, F.; Ferris, S.; Möbius, H.J. Memantine in Moderate-to-Severe Alzheimer’s Disease. N. Engl. J. Med. 2003, 348, 1333–1341. [Google Scholar] [CrossRef]
  8. Robinson, D.M.; Keating, G.M. Memantine. Drugs 2006, 66, 1515–1534. [Google Scholar] [CrossRef]
  9. Eyjolfsdottir, H.; Eriksdotter, M.; Linderoth, B.; Lind, G.; Juliusson, B.; Kusk, P.; Almkvist, O.; Andreasen, N.; Blennow, K.; Ferreira, D.; et al. Targeted Delivery of Nerve Growth Factor to the Cholinergic Basal Forebrain of Alzheimer’s Disease Patients: Application of a Second-Generation Encapsulated Cell Biodelivery Device. Alzheimers Res. Ther. 2016, 8, 30. [Google Scholar] [CrossRef]
  10. Birks, J.S.; Harvey, R.J. Donepezil for Dementia Due to Alzheimer’s Disease. Cochrane Database Syst. Rev. 2018, 2018, CD001190. [Google Scholar] [CrossRef]
  11. Kennedy, M.E.; Stamford, A.W.; Chen, X.; Cox, K.; Cumming, J.N.; Dockendorf, M.F.; Egan, M.; Ereshefsky, L.; Hodgson, R.A.; Hyde, L.A.; et al. The BACE1 Inhibitor Verubecestat (MK-8931) Reduces CNS β-Amyloid in Animal Models and in Alzheimer’s Disease Patients. Sci. Transl. Med. 2016, 8, 363ra150. [Google Scholar] [CrossRef]
  12. Novak, P.; Schmidt, R.; Kontsekova, E.; Zilka, N.; Kovacech, B.; Skrabana, R.; Vince-Kazmerova, Z.; Katina, S.; Fialova, L.; Prcina, M.; et al. Safety and Immunogenicity of the Tau Vaccine AADvac1 in Patients with Alzheimer’s Disease: A Randomised, Double-Blind, Placebo-Controlled, Phase 1 Trial. Lancet Neurol. 2017, 16, 123–134. [Google Scholar] [CrossRef]
  13. Yamada, K.; Tanaka, T.; Han, D.; Senzaki, K.; Kameyama, T.; Nabeshima, T. Protective Effects of Idebenone and α-Tocopherol on β-Amyloid-(1–42)-Induced Learning and Memory Deficits in Rats: Implication of Oxidative Stress in β-Amyloid-Induced Neurotoxicity in Vivo. Eur. J. Neurosci. 1999, 11, 83–90. [Google Scholar] [CrossRef] [PubMed]
  14. Ohkura, T.; Isse, K.; Akazawa, K.; Hamamoto, M.; Yaoi, Y.; Hagino, N. Evaluation of Estrogen Treatment in Female Patients with Dementia of the Alzheimer Type. Endocr. J. 1994, 41, 361–371. [Google Scholar] [CrossRef]
  15. Ohkura, T.; Isse, K.; Akazawa, K.; Hamamoto, M.; Yaoi, Y.; Hagino, N. Long-Term Estrogen Replacement Therapy in Female Patients with Dementia of the Alzheimer Type: 7 Case Reports. Dement. Geriatr. Cogn. Disord. 1995, 6, 99–107. [Google Scholar] [CrossRef] [PubMed]
  16. Baranowska-Wójcik, E.; Szwajgier, D.; Winiarska-Mieczan, A. Honey as the Potential Natural Source of Cholinesterase Inhibitors in Alzheimer’s Disease. Plant Foods Hum. Nutr. 2020, 75, 30–32. [Google Scholar] [CrossRef] [PubMed]
  17. Azman, K.F.; Zakaria, R. Honey as an Antioxidant Therapy to Reduce Cognitive Ageing. Iran. J. Basic Med. Sci. 2019, 22, 1368–1377. [Google Scholar] [CrossRef] [PubMed]
  18. Nordin, A.; Saim, A.B.; Idrus, R.B.H. Honey Ameliorate Negative Effects in Neurodegenerative Diseases: An Evidence-Based Review. Sains Malays. 2021, 50, 791–801. [Google Scholar] [CrossRef]
  19. Kuns, B.; Rosani, A.; Varghese, D. Memantine. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  20. Rossom, R.; Adityanjee; Dysken, M. Efficacy and Tolerability of Memantine in the Treatment of Dementia. Am. J. Geriatr. Pharmacother. 2004, 2, 303–312. [Google Scholar] [CrossRef]
  21. Li, S.; Liu, L.; Selkoe, D. Verubecestat for Prodromal Alzheimer’s Disease. N. Engl. J. Med. 2019, 381, 388–389. [Google Scholar] [CrossRef]
  22. Sperling, R.; Salloway, S.; Brooks, D.J.; Tampieri, D.; Barakos, J.; Fox, N.C.; Raskind, M.; Sabbagh, M.; Honig, L.S.; Porsteinsson, A.P.; et al. Amyloid-Related Imaging Abnormalities in Patients with Alzheimer’s Disease Treated with Bapineuzumab: A Retrospective Analysis. Lancet Neurol. 2012, 11, 241–249. [Google Scholar] [CrossRef]
  23. Farina, N.; Llewellyn, D.; Isaac, M.G.E.K.N.; Tabet, N. Vitamin E for Alzheimer’s Dementia and Mild Cognitive Impairment. Cochrane Database Syst. Rev. 2017, 4, CD002854. [Google Scholar] [CrossRef]
  24. Farkas, S.; Szabó, A.; Hegyi, A.E.; Török, B.; Fazekas, C.L.; Ernszt, D.; Kovács, T.; Zelena, D. Estradiol and Estrogen-like Alternative Therapies in Use: The Importance of the Selective and Non-Classical Actions. Biomedicines 2022, 10, 861. [Google Scholar] [CrossRef] [PubMed]
  25. Artero, A.; Tarín, J.J.; Cano, A. The Adverse Effects of Estrogen and Selective Estrogen Receptor Modulators on Hemostasis and Thrombosis. Semin. Thromb. Hemost. 2012, 38, 797–807. [Google Scholar] [CrossRef]
  26. Bauer, L.; Kohlich, A.; Hirschwehr, R.; Siemann, U.; Ebner, H.; Scheiner, O.; Kraft, D.; Ebner, C. Food Allergy to Honey: Pollen or Bee Products?: Characterization of Allergenic Proteins in Honey by Means of Immunoblotting. J. Allergy Clin. Immunol. 1996, 97, 65–73. [Google Scholar] [CrossRef] [PubMed]
  27. Di Costanzo, M.; De Paulis, N.; Peveri, S.; Montagni, M.; Canani, R.B.; Biasucci, G. Anaphylaxis Caused by Artisanal Honey in a Child: A Case Report. J. Med. Case Reports 2021, 15, 235. [Google Scholar] [CrossRef] [PubMed]
  28. Aguiar, R.; Duarte, F.C.; Mendes, A.; Bartolomé, B.; Barbosa, M.P. Anaphylaxis Caused by Honey: A Case Report. Asia Pac. Allergy 2017, 7, 48–50. [Google Scholar] [CrossRef] [PubMed]
  29. Al-Rahbi, B.; Zakaria, R.; Othman, Z.; Hassan, A.; Ahmad, A.H. Protective Effects of Tualang Honey against Oxidative Stress and Anxiety-Like Behaviour in Stressed Ovariectomized Rats. Int. Sch. Res. Not. 2014, 2014, e521065. [Google Scholar] [CrossRef]
  30. Sairazi, N.S.M.; Sirajudeen, K.N.S.; Asari, M.A.; Mummedy, S.; Muzaimi, M.; Sulaiman, S.A. Effect of Tualang Honey against KA-Induced Oxidative Stress and Neurodegeneration in the Cortex of Rats. BMC Complement. Altern. Med. 2017, 17, 31. [Google Scholar] [CrossRef]
  31. Shafin, N.; Othman, Z.; Zakaria, R.; Hussain, N.H.N. Tualang Honey Supplementation Reduces Blood Oxidative Stress Levels/Activities in Postmenopausal Women. Int. Sch. Res. Not. 2014, 2014, e364836. [Google Scholar] [CrossRef]
  32. Candiracci, M.; Piatti, E.; Dominguez-Barragán, M.; García-Antrás, D.; Morgado, B.; Ruano, D.; Gutiérrez, J.F.; Parrado, J.; Castaño, A. Anti-Inflammatory Activity of a Honey Flavonoid Extract on Lipopolysaccharide-Activated N13 Microglial Cells. J. Agric. Food Chem. 2012, 60, 12304–12311. [Google Scholar] [CrossRef]
  33. Sairazi, N.S.M.; Sirajudeen, K.N.S.; Muzaimi, M.; Mummedy, S.; Asari, M.A.; Sulaiman, S.A. Tualang Honey Reduced Neuroinflammation and Caspase-3 Activity in Rat Brain after Kainic Acid-Induced Status Epilepticus. Evid. Based Complement. Alternat. Med. 2018, 2018, e7287820. [Google Scholar] [CrossRef]
  34. Arshad, N.A.; Lin, T.S.; Yahaya, M.F. Stingless Bee Honey Reduces Anxiety and Improves Memory of the Metabolic Disease-Induced Rats. CNS Neurol. Disord.-Drug Targets-CNS Neurol. Disord. 2020, 19, 115–126. [Google Scholar] [CrossRef] [PubMed]
  35. Saxena, A.K.; Phyu, H.P.; Al-Ani, I.M.; Talib, N.A. Potential Protective Effect of Honey against Chronic Cerebral Hypoperfusion-Induced Neurodegeneration in Rats. J. Anat. Soc. India 2014, 63, 151–155. [Google Scholar] [CrossRef]
  36. Zaidi, H.; Ouchemoukh, S.; Amessis-Ouchemoukh, N.; Debbache, N.; Pacheco, R.; Serralheiro, M.L.; Araujo, M.E. Biological Properties of Phenolic Compound Extracts in Selected Algerian Honeys—The Inhibition of Acetylcholinesterase and α-Glucosidase Activities. Eur. J. Integr. Med. 2019, 25, 77–84. [Google Scholar] [CrossRef]
  37. Al-Himyari, F.A. P1-241: The Use of Honey as a Natural Preventive Therapy of Cognitive Decline and Dementia in the Middle East. Alzheimers Dement. 2009, 5, P247. [Google Scholar] [CrossRef]
  38. Shafin, N.; Zakaria, R.; Othman, Z.; Nik, N.H. Improved Blood Oxidative Status Is Not Associated with Better Memory Performance in Postmenopausal Women Receiving Tualang Honey Supplementation. J. Biochem. Pharmacol. Res. 2014, 2, 110–116. [Google Scholar]
  39. Behl, C. Alzheimer’s Disease and Oxidative Stress: Implications for Novel Therapeutic Approaches. Prog. Neurobiol. 1999, 57, 301–323. [Google Scholar] [CrossRef] [PubMed]
  40. Schönheit, B.; Zarski, R.; Ohm, T.G. Spatial and Temporal Relationships between Plaques and Tangles in Alzheimer-Pathology. Neurobiol. Aging 2004, 25, 697–711. [Google Scholar] [CrossRef]
  41. Vossel, K.A.; Zhang, K.; Brodbeck, J.; Daub, A.C.; Sharma, P.; Finkbeiner, S.; Cui, B.; Mucke, L. Tau Reduction Prevents Aβ-Induced Defects in Axonal Transport. Science 2010, 330, 198. [Google Scholar] [CrossRef]
  42. Braak, H.; Del Tredici, K. Evolutional Aspects of Alzheimer’s Disease Pathogenesis. J. Alzheimers Dis. 2013, 33, S155–S161. [Google Scholar] [CrossRef]
  43. Ellmerich, S.; Taylor, G.W.; Richardson, C.D.; Minett, T.; Schmidt, A.F.; Brayne, C.; Matthews, F.E.; Ince, P.G.; Wharton, S.B.; Pepys, M.B.; et al. Dementia in the Older Population Is Associated with Neocortex Content of Serum Amyloid P Component. Brain Commun. 2021, 3, fcab225. [Google Scholar] [CrossRef]
  44. Crawford, J.R.; Bjorklund, N.L.; Taglialatela, G.; Gomer, R.H. Brain Serum Amyloid P Levels Are Reduced in Individuals That Lack Dementia While Having Alzheimer’s Disease Neuropathology. Neurochem. Res. 2012, 37, 795–801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Kuchibhotla, K.V.; Wegmann, S.; Kopeikina, K.J.; Hawkes, J.; Rudinskiy, N.; Andermann, M.L.; Spires-Jones, T.L.; Bacskai, B.J.; Hyman, B.T. Neurofibrillary Tangle-Bearing Neurons Are Functionally Integrated in Cortical Circuits in Vivo. Proc. Natl. Acad. Sci. USA 2014, 111, 510–514. [Google Scholar] [CrossRef] [PubMed]
  46. Haroutunian, V.; Schnaider-Beeri, M.; Schmeidler, J.; Wysocki, M.; Purohit, D.P.; Perl, D.P.; Libow, L.S.; Lesser, G.T.; Maroukian, M.; Grossman, H.T. Role of the Neuropathology of Alzheimer Disease in Dementia in the Oldest-Old. Arch. Neurol. 2008, 65, 1211–1217. [Google Scholar] [CrossRef]
  47. Li, T.; Braunstein, K.E.; Zhang, J.; Lau, A.; Sibener, L.; Deeble, C.; Wong, P.C. The Neuritic Plaque Facilitates Pathological Conversion of Tau in an Alzheimer’s Disease Mouse Model. Nat. Commun. 2016, 7, 12082. [Google Scholar] [CrossRef] [PubMed]
  48. Khan, U.A.; Liu, L.; Provenzano, F.A.; Berman, D.E.; Profaci, C.P.; Sloan, R.; Mayeux, R.; Duff, K.E.; Small, S.A. Molecular Drivers and Cortical Spread of Lateral Entorhinal Cortex Dysfunction in Preclinical Alzheimer’s Disease. Nat. Neurosci. 2014, 17, 304–311. [Google Scholar] [CrossRef]
  49. Nelson, P.T.; Abner, E.L.; Schmitt, F.A.; Kryscio, R.J.; Jicha, G.A.; Santacruz, K.; Smith, C.D.; Patel, E.; Markesbery, W.R. Brains With Medial Temporal Lobe Neurofibrillary Tangles But No Neuritic Amyloid Plaques Are a Diagnostic Dilemma But May Have Pathogenetic Aspects Distinct From Alzheimer Disease. J. Neuropathol. Exp. Neurol. 2009, 68, 774–784. [Google Scholar] [CrossRef]
  50. Li, L.; Jiang, Y.; Wang, J.-Z.; Liu, R.; Wang, X. Tau Ubiquitination in Alzheimer’s Disease. Front. Neurol. 2022, 12, 786353. [Google Scholar] [CrossRef]
  51. Rodríguez-Martín, T.; Cuchillo-Ibáñez, I.; Noble, W.; Nyenya, F.; Anderton, B.H.; Hanger, D.P. Tau Phosphorylation Affects Its Axonal Transport and Degradation. Neurobiol. Aging 2013, 34, 2146–2157. [Google Scholar] [CrossRef]
  52. Yoshiyama, Y.; Higuchi, M.; Zhang, B.; Huang, S.-M.; Iwata, N.; Saido, T.C.; Maeda, J.; Suhara, T.; Trojanowski, J.Q.; Lee, V.M.-Y. Synapse Loss and Microglial Activation Precede Tangles in a P301S Tauopathy Mouse Model. Neuron 2007, 53, 337–351. [Google Scholar] [CrossRef]
  53. Arnaud, L.; Robakis, N.K.; Figueiredo-Pereira, M.E. It May Take Inflammation, Phosphorylation and Ubiquitination to ‘Tangle’ in Alzheimer’s Disease. Neurodegener. Dis. 2006, 3, 313–319. [Google Scholar] [CrossRef]
  54. Lennol, M.P.; Sánchez-Domínguez, I.; Cuchillo-Ibañez, I.; Camporesi, E.; Brinkmalm, G.; Alcolea, D.; Fortea, J.; Lleó, A.; Soria, G.; Aguado, F.; et al. Apolipoprotein E Imbalance in the Cerebrospinal Fluid of Alzheimer’s Disease Patients. Alzheimers Res. Ther. 2022, 14, 161. [Google Scholar] [CrossRef]
  55. Corder, E.H.; Saunders, A.M.; Strittmatter, W.J.; Schmechel, D.E.; Gaskell, P.C.; Small, G.W.; Roses, A.D.; Haines, J.L.; Pericak-Vance, M.A. Gene Dose of Apolipoprotein E Type 4 Allele and the Risk of Alzheimer’s Disease in Late Onset Families. Science 1993, 261, 921–923. [Google Scholar] [CrossRef]
  56. Saunders, A.M.; Strittmatter, W.J.; Schmechel, D.; George-Hyslop, P.H.S.; Pericak-Vance, M.A.; Joo, S.H.; Rosi, B.L.; Gusella, J.F.; Crapper-MacLachlan, D.R.; Alberts, M.J.; et al. Association of Apolipoprotein E Allele Ε4 with Late-onset Familial and Sporadic Alzheimer’s Disease. Neurology 1993, 43, 1467–1472. [Google Scholar] [CrossRef]
  57. Miyata, M.; Smith, J.D. Apolipoprotein E Allele–Specific Antioxidant Activity and Effects on Cytotoxicity by Oxidative Insults and β–Amyloid Peptides. Nat. Genet. 1996, 14, 55–61. [Google Scholar] [CrossRef]
  58. Butterfield, D.A.; Mattson, M.P. Apolipoprotein E and Oxidative Stress in Brain with Relevance to Alzheimer’s Disease. Neurobiol. Dis. 2020, 138, 104795. [Google Scholar] [CrossRef]
  59. Gralle, M.; Ferreira, S.T. Structure and Functions of the Human Amyloid Precursor Protein: The Whole Is More than the Sum of Its Parts. Prog. Neurobiol. 2007, 82, 11–32. [Google Scholar] [CrossRef]
  60. Rovelet-Lecrux, A.; Hannequin, D.; Raux, G.; Meur, N.L.; Laquerrière, A.; Vital, A.; Dumanchin, C.; Feuillette, S.; Brice, A.; Vercelletto, M.; et al. APP Locus Duplication Causes Autosomal Dominant Early-Onset Alzheimer Disease with Cerebral Amyloid Angiopathy. Nat. Genet. 2006, 38, 24–26. [Google Scholar] [CrossRef]
  61. Sleegers, K.; Brouwers, N.; Gijselinck, I.; Theuns, J.; Goossens, D.; Wauters, J.; Del-Favero, J.; Cruts, M.; van Duijn, C.M.; Broeckhoven, C.V. APP Duplication Is Sufficient to Cause Early Onset Alzheimer’s Dementia with Cerebral Amyloid Angiopathy. Brain 2006, 129, 2977–2983. [Google Scholar] [CrossRef]
  62. Wiseman, F.K.; Al-Janabi, T.; Hardy, J.; Karmiloff-Smith, A.; Nizetic, D.; Tybulewicz, V.L.J.; Fisher, E.M.C.; Strydom, A. A Genetic Cause of Alzheimer Disease: Mechanistic Insights from Down Syndrome. Nat. Rev. Neurosci. 2015, 16, 564–574. [Google Scholar] [CrossRef]
  63. Zhao, J.; Fu, Y.; Yasvoina, M.; Shao, P.; Hitt, B.; O’Connor, T.; Logan, S.; Maus, E.; Citron, M.; Berry, R.; et al. β-Site Amyloid Precursor Protein Cleaving Enzyme 1 Levels Become Elevated in Neurons around Amyloid Plaques: Implications for Alzheimer’s Disease Pathogenesis. J. Neurosci. 2007, 27, 3639–3649. [Google Scholar] [CrossRef]
  64. Cai, Y.; An, S.S.A.; Kim, S. Mutations in Presenilin 2 and Its Implications in Alzheimer’s Disease and Other Dementia-Associated Disorders. Clin. Interv. Aging 2015, 10, 1163–1172. [Google Scholar] [CrossRef] [Green Version]
  65. Hampel, H.; Hardy, J.; Blennow, K.; Chen, C.; Perry, G.; Kim, S.H.; Villemagne, V.L.; Aisen, P.; Vendruscolo, M.; Iwatsubo, T.; et al. The Amyloid-β Pathway in Alzheimer’s Disease. Mol. Psychiatry 2021, 26, 5481–5503. [Google Scholar] [CrossRef]
  66. Wang, M.; Jing, T.; Wang, X.; Yao, D. Beta-Secretase/BACE1 Promotes APP Endocytosis and Processing in the Endosomes and on Cell Membrane. Neurosci. Lett. 2018, 685, 63–67. [Google Scholar] [CrossRef]
  67. Siegel, G.; Gerber, H.; Koch, P.; Bruestle, O.; Fraering, P.C.; Rajendran, L. The Alzheimer’s Disease γ-Secretase Generates Higher 42:40 Ratios for β-Amyloid Than for P3 Peptides. Cell Rep. 2017, 19, 1967–1976. [Google Scholar] [CrossRef]
  68. Tyler, S.J.; Dawbarn, D.; Wilcock, G.K.; Allen, S.J. α- and β-Secretase: Profound Changes in Alzheimer’s Disease. Biochem. Biophys. Res. Commun. 2002, 299, 373–376. [Google Scholar] [CrossRef]
  69. Scahill, R.I.; Schott, J.M.; Stevens, J.M.; Rossor, M.N.; Fox, N.C. Mapping the Evolution of Regional Atrophy in Alzheimer’s Disease: Unbiased Analysis of Fluid-Registered Serial MRI. Proc. Natl. Acad. Sci. USA 2002, 99, 4703–4707. [Google Scholar] [CrossRef]
  70. Chetelat, G.A.; Baron, J.-C. Early Diagnosis of Alzheimer’s Disease: Contribution of Structural Neuroimaging. NeuroImage 2003, 18, 525–541. [Google Scholar] [CrossRef]
  71. Scheff, S.W.; Price, D.A.; Schmitt, F.A.; Mufson, E.J. Hippocampal Synaptic Loss in Early Alzheimer’s Disease and Mild Cognitive Impairment. Neurobiol. Aging 2006, 27, 1372–1384. [Google Scholar] [CrossRef]
  72. Bell, K.F.S.; Ducatenzeiler, A.; Ribeiro-da-Silva, A.; Duff, K.; Bennett, D.A.; Cuello, A.C. The Amyloid Pathology Progresses in a Neurotransmitter-Specific Manner. Neurobiol. Aging 2006, 27, 1644–1657. [Google Scholar] [CrossRef]
  73. Jack, C.R., Jr.; Lowe, V.J.; Weigand, S.D.; Wiste, H.J.; Senjem, M.L.; Knopman, D.S.; Shiung, M.M.; Gunter, J.L.; Boeve, B.F.; Kemp, B.J.; et al. Serial PIB and MRI in Normal, Mild Cognitive Impairment and Alzheimer’s Disease: Implications for Sequence of Pathological Events in Alzheimer’s Disease. Brain 2009, 132, 1355–1365. [Google Scholar] [CrossRef]
  74. Fitzjohn, S.M.; Morton, R.A.; Kuenzi, F.; Rosahl, T.W.; Shearman, M.; Lewis, H.; Smith, D.; Reynolds, D.S.; Davies, C.H.; Collingridge, G.L.; et al. Age-Related Impairment of Synaptic Transmission But Normal Long-Term Potentiation in Transgenic Mice That Overexpress the Human APP695SWE Mutant Form of Amyloid Precursor Protein. J. Neurosci. 2001, 21, 4691–4698. [Google Scholar] [CrossRef] [PubMed]
  75. Jacobsen, J.S.; Wu, C.-C.; Redwine, J.M.; Comery, T.A.; Arias, R.; Bowlby, M.; Martone, R.; Morrison, J.H.; Pangalos, M.N.; Reinhart, P.H.; et al. Early-Onset Behavioral and Synaptic Deficits in a Mouse Model of Alzheimer’s Disease. Proc. Natl. Acad. Sci. USA 2006, 103, 5161–5166. [Google Scholar] [CrossRef] [PubMed]
  76. Itoh, A.; Akaike, T.; Sokabe, M.; Nitta, A.; Iida, R.; Olariu, A.; Yamada, K.; Nabeshima, T. Impairments of Long-Term Potentiation in Hippocampal Slices of β-Amyloid-Infused Rats. Eur. J. Pharmacol. 1999, 382, 167–175. [Google Scholar] [CrossRef] [PubMed]
  77. Jung, J.H.; An, K.; Kwon, O.B.; Kim, H.; Kim, J.-H. Pathway-Specific Alteration of Synaptic Plasticity in Tg2576 Mice. Mol. Cells 2011, 32, 197–201. [Google Scholar] [CrossRef] [PubMed]
  78. Witton, J.; Brown, J.T.; Jones, M.W.; Randall, A.D. Altered Synaptic Plasticity in the Mossy Fibre Pathway of Transgenic Mice Expressing Mutant Amyloid Precursor Protein. Mol. Brain 2010, 3, 32. [Google Scholar] [CrossRef] [PubMed]
  79. Ansari, M.A.; Scheff, S.W. Oxidative Stress in the Progression of Alzheimer Disease in the Frontal Cortex. J. Neuropathol. Exp. Neurol. 2010, 69, 155–167. [Google Scholar] [CrossRef]
  80. Butterfield, D.A.; Reed, T.T.; Perluigi, M.; De Marco, C.; Coccia, R.; Keller, J.N.; Markesbery, W.R.; Sultana, R. Elevated Levels of 3-Nitrotyrosine in Brain from Subjects with Amnestic Mild Cognitive Impairment: Implications for the Role of Nitration in the Progression of Alzheimer’s Disease. Brain Res. 2007, 1148, 243–248. [Google Scholar] [CrossRef]
  81. Bradley-Whitman, M.A.; Lovell, M.A. Biomarkers of Lipid Peroxidation in Alzheimer Disease (AD): An Update. Arch. Toxicol. 2015, 89, 1035–1044. [Google Scholar] [CrossRef]
  82. Kinney, J.W.; Bemiller, S.M.; Murtishaw, A.S.; Leisgang, A.M.; Salazar, A.M.; Lamb, B.T. Inflammation as a Central Mechanism in Alzheimer’s Disease. Alzheimers Dement. Transl. Res. Clin. Interv. 2018, 4, 575–590. [Google Scholar] [CrossRef]
  83. Lyra e Silva, N.M.; Gonçalves, R.A.; Pascoal, T.A.; Lima-Filho, R.A.S.; de Paula França Resende, E.; Vieira, E.L.M.; Teixeira, A.L.; de Souza, L.C.; Peny, J.A.; Fortuna, J.T.S.; et al. Pro-Inflammatory Interleukin-6 Signaling Links Cognitive Impairments and Peripheral Metabolic Alterations in Alzheimer’s Disease. Transl. Psychiatry 2021, 11, 251. [Google Scholar] [CrossRef]
  84. González-Reyes, R.E.; Nava-Mesa, M.O.; Vargas-Sánchez, K.; Ariza-Salamanca, D.; Mora-Muñoz, L. Involvement of Astrocytes in Alzheimer’s Disease from a Neuroinflammatory and Oxidative Stress Perspective. Front. Mol. Neurosci. 2017, 10, 427. [Google Scholar] [CrossRef] [Green Version]
  85. Granic, I.; Dolga, A.M.; Nijholt, I.M.; van Dijk, G.; Eisel, U.L.M. Inflammation and NF-ΚB in Alzheimer’s Disease and Diabetes. J. Alzheimers Dis. 2009, 16, 809–821. [Google Scholar] [CrossRef]
  86. Onyango, I.G.; Jauregui, G.V.; Čarná, M.; Bennett, J.P.; Stokin, G.B. Neuroinflammation in Alzheimer’s Disease. Biomedicines 2021, 9, 524. [Google Scholar] [CrossRef]
  87. Su, F.; Bai, F.; Zhang, Z. Inflammatory Cytokines and Alzheimer’s Disease: A Review from the Perspective of Genetic Polymorphisms. Neurosci. Bull. 2016, 32, 469–480. [Google Scholar] [CrossRef]
  88. Gadani, S.P.; Cronk, J.C.; Norris, G.T.; Kipnis, J. IL-4 in the Brain: A Cytokine To Remember. J. Immunol. 2012, 189, 4213–4219. [Google Scholar] [CrossRef]
  89. Wang, P.; Wu, P.; Siegel, M.I.; Egan, R.W.; Billah, M.M. Interleukin (IL)-10 Inhibits Nuclear Factor KB (NFĸB) Activation in Human Monocytes: IL-10 AND IL-4 SUPPRESS CYTOKINE SYNTHESIS BY DIFFERENT MECHANISMS (∗). J. Biol. Chem. 1995, 270, 9558–9563. [Google Scholar] [CrossRef]
  90. Friedberg, J.S.; Aytan, N.; Cherry, J.D.; Xia, W.; Standring, O.J.; Alvarez, V.E.; Nicks, R.; Svirsky, S.; Meng, G.; Jun, G.; et al. Associations between Brain Inflammatory Profiles and Human Neuropathology Are Altered Based on Apolipoprotein E Ε4 Genotype. Sci. Rep. 2020, 10, 2924. [Google Scholar] [CrossRef]
  91. Tai, L.M.; Ghura, S.; Koster, K.P.; Liakaite, V.; Maienschein-Cline, M.; Kanabar, P.; Collins, N.; Ben-Aissa, M.; Lei, A.Z.; Bahroos, N.; et al. APOE-Modulated Aβ-Induced Neuroinflammation in Alzheimer’s Disease: Current Landscape, Novel Data, and Future Perspective. J. Neurochem. 2015, 133, 465–488. [Google Scholar] [CrossRef]
  92. Jessen, F.; Feyen, L.; Freymann, K.; Tepest, R.; Maier, W.; Heun, R.; Schild, H.-H.; Scheef, L. Volume Reduction of the Entorhinal Cortex in Subjective Memory Impairment. Neurobiol. Aging 2006, 27, 1751–1756. [Google Scholar] [CrossRef]
  93. Ajibola, A.; Chamunorwa, J.P.; Erlwanger, K.H. Nutraceutical Values of Natural Honey and Its Contribution to Human Health and Wealth. Nutr. Metab. 2012, 9, 61. [Google Scholar] [CrossRef]
  94. Kreider, R.B.; Rasmussen, C.J.; Lancaster, S.L.; Kerksick, C.; Greenwood, M. Honey: An Alternative Sports Gel. Strength Cond. J. 2002, 24, 50–51. [Google Scholar] [CrossRef]
  95. Santos-Buelga, C.; González-Paramás, A.M. Chemical Composition of Honey. In Bee Products—Chemical and Biological Properties; Alvarez-Suarez, J.M., Ed.; Springer International Publishing: Cham, Switzerland, 2017; pp. 43–82. ISBN 978-3-319-59689-1. [Google Scholar]
  96. Bogdanov, S.; Jurendic, T.; Sieber, R.; Gallmann, P. Honey for Nutrition and Health: A Review. J. Am. Coll. Nutr. 2008, 27, 677–689. [Google Scholar] [CrossRef] [PubMed]
  97. Olas, B. Honey and Its Phenolic Compounds as an Effective Natural Medicine for Cardiovascular Diseases in Humans? Nutrients 2020, 12, 283. [Google Scholar] [CrossRef]
  98. Cheung, Y.; Meenu, M.; Yu, X.; Xu, B. Phenolic Acids and Flavonoids Profiles of Commercial Honey from Different Floral Sources and Geographic Sources. Int. J. Food Prop. 2019, 22, 290–308. [Google Scholar] [CrossRef]
  99. Sousa, J.M.; de Souza, E.L.; Marques, G.; Meireles, B.; de Magalhães Cordeiro, Â.T.; Gullón, B.; Pintado, M.M.; Magnani, M. Polyphenolic Profile and Antioxidant and Antibacterial Activities of Monofloral Honeys Produced by Meliponini in the Brazilian Semiarid Region. Food Res. Int. 2016, 84, 61–68. [Google Scholar] [CrossRef]
  100. Ranneh, Y.; Ali, F.; Zarei, M.; Akim, A.M.; Hamid, H.A.; Khazaai, H. Malaysian Stingless Bee and Tualang Honeys: A Comparative Characterization of Total Antioxidant Capacity and Phenolic Profile Using Liquid Chromatography-Mass Spectrometry. LWT 2018, 89, 1–9. [Google Scholar] [CrossRef]
  101. Kek, S.P.; Chin, N.L.; Yusof, Y.A.; Tan, S.W.; Chua, L.S. Total Phenolic Contents and Colour Intensity of Malaysian Honeys from the Apis Spp. and Trigona Spp. Bees. Agric. Agric. Sci. Procedia 2014, 2, 150–155. [Google Scholar] [CrossRef]
  102. Moniruzzaman, M.; Khalil, M.I.; Sulaiman, S.A.; Gan, S.H. Physicochemical and Antioxidant Properties of Malaysian Honeys Produced by Apis Cerana, Apis Dorsata and Apis Mellifera. BMC Complement. Altern. Med. 2013, 13, 43. [Google Scholar] [CrossRef]
  103. Kishore, R.K.; Halim, A.S.; Syazana, M.S.N.; Sirajudeen, K.N.S. Tualang Honey Has Higher Phenolic Content and Greater Radical Scavenging Activity Compared with Other Honey Sources. Nutr. Res. 2011, 31, 322–325. [Google Scholar] [CrossRef]
  104. Putteeraj, M.; Lim, W.L.; Teoh, S.L.; Yahaya, M.F. Flavonoids and Its Neuroprotective Effects on Brain Ischemia and Neurodegenerative Diseases. Curr. Drug Targets 2018, 19, 1710–1720. [Google Scholar] [CrossRef]
  105. Mohd Kamal, D.A.; Ibrahim, S.F.; Kamal, H.; Kashim, M.I.A.M.; Mokhtar, M.H. Physicochemical and Medicinal Properties of Tualang, Gelam and Kelulut Honeys: A Comprehensive Review. Nutrients 2021, 13, 197. [Google Scholar] [CrossRef]
  106. Floyd, R.A. Neuroinflammatory Processes Are Important in Neurodegenerative Diseases: An Hypothesis to Explain the Increased Formation of Reactive Oxygen and Nitrogen Species as Major Factors Involved in Neurodegenerative Disease Development. Free Radic. Biol. Med. 1999, 26, 1346–1355. [Google Scholar] [CrossRef]
  107. Calabrese, V.; Bates, T.E.; Stella, A.M.G. NO Synthase and NO-Dependent Signal Pathways in Brain Aging and Neurodegenerative Disorders: The Role of Oxidant/Antioxidant Balance. Neurochem. Res. 2000, 25, 1315–1341. [Google Scholar] [CrossRef]
  108. Pluta, R.; Januszewski, S.; Czuczwar, S.J. Myricetin as a Promising Molecule for the Treatment of Post-Ischemic Brain Neurodegeneration. Nutrients 2021, 13, 342. [Google Scholar] [CrossRef]
  109. Ramezani, M.; Darbandi, N.; Khodagholi, F.; Hashemi, A. Myricetin Protects Hippocampal CA3 Pyramidal Neurons and Improves Learning and Memory Impairments in Rats with Alzheimer’s Disease. Neural Regen. Res. 2016, 11, 1976–1980. [Google Scholar] [CrossRef]
  110. Wang, B.; Zhong, Y.; Gao, C.; Li, J. Myricetin Ameliorates Scopolamine-Induced Memory Impairment in Mice via Inhibiting Acetylcholinesterase and down-Regulating Brain Iron. Biochem. Biophys. Res. Commun. 2017, 490, 336–342. [Google Scholar] [CrossRef]
  111. Shimmyo, Y.; Kihara, T.; Akaike, A.; Niidome, T.; Sugimoto, H. Multifunction of Myricetin on Aβ: Neuroprotection via a Conformational Change of Aβ and Reduction of Aβ via the Interference of Secretases. J. Neurosci. Res. 2008, 86, 368–377. [Google Scholar] [CrossRef]
  112. Wang, H.; Wang, H.; Cheng, H.; Che, Z. Ameliorating Effect of Luteolin on Memory Impairment in an Alzheimer’s Disease Model. Mol. Med. Rep. 2016, 13, 4215–4220. [Google Scholar] [CrossRef]
  113. Fu, X.; Zhang, J.; Guo, L.; Xu, Y.; Sun, L.; Wang, S.; Feng, Y.; Gou, L.; Zhang, L.; Liu, Y. Protective Role of Luteolin against Cognitive Dysfunction Induced by Chronic Cerebral Hypoperfusion in Rats. Pharmacol. Biochem. Behav. 2014, 126, 122–130. [Google Scholar] [CrossRef]
  114. Jang, S.; Dilger, R.N.; Johnson, R.W. Luteolin Inhibits Microglia and Alters Hippocampal-Dependent Spatial Working Memory in Aged Mice. J. Nutr. 2010, 140, 1892–1898. [Google Scholar] [CrossRef]
  115. Zhou, T.; Liu, L.; Wang, Q.; Gao, Y. Naringenin Alleviates Cognition Deficits in High-Fat Diet-Fed SAMP8 Mice. J. Food Biochem. 2020, 44, e13375. [Google Scholar] [CrossRef]
  116. Haider, S.; Liaquat, L.; Ahmad, S.; Batool, Z.; Siddiqui, R.A.; Tabassum, S.; Shahzad, S.; Rafiq, S.; Naz, N. Naringenin Protects AlCl3/D-Galactose Induced Neurotoxicity in Rat Model of AD via Attenuation of Acetylcholinesterase Levels and Inhibition of Oxidative Stress. PLOS ONE 2020, 15, e0227631. [Google Scholar] [CrossRef] [PubMed]
  117. Ghofrani, S.; Joghataei, M.-T.; Mohseni, S.; Baluchnejadmojarad, T.; Bagheri, M.; Khamse, S.; Roghani, M. Naringenin Improves Learning and Memory in an Alzheimer’s Disease Rat Model: Insights into the Underlying Mechanisms. Eur. J. Pharmacol. 2015, 764, 195–201. [Google Scholar] [CrossRef] [PubMed]
  118. Khan, M.B.; Khan, M.M.; Khan, A.; Ahmed, E.; Ishrat, T.; Tabassum, R.; Vaibhav, K.; Ahmad, A.; Islam, F. Naringenin Ameliorates Alzheimer’s Disease (AD)-Type Neurodegeneration with Cognitive Impairment (AD-TNDCI) Caused by the Intracerebroventricular-Streptozotocin in Rat Model. Neurochem. Int. 2012, 61, 1081–1093. [Google Scholar] [CrossRef] [PubMed]
  119. Meng, X.; Fu, M.; Wang, S.; Chen, W.; Wang, J.; Zhang, N. Naringin Ameliorates Memory Deficits and Exerts Neuroprotective Effects in a Mouse Model of Alzheimer’s Disease by Regulating Multiple Metabolic Pathways. Mol. Med. Rep. 2021, 23, 332. [Google Scholar] [CrossRef]
  120. Tongjaroenbuangam, W.; Ruksee, N.; Chantiratikul, P.; Pakdeenarong, N.; Kongbuntad, W.; Govitrapong, P. Neuroprotective Effects of Quercetin, Rutin and Okra (Abelmoschus Esculentus Linn.) in Dexamethasone-Treated Mice. Neurochem. Int. 2011, 59, 677–685. [Google Scholar] [CrossRef]
  121. Ashrafpour, M.; Parsaei, S.; Sepehri, H. Quercetin Improved Spatial Memory Dysfunctions in Rat Model of Intracerebroventricular Streptozotocin-Induced Sporadic Alzheimer’sdisease. Natl. J. Physiol. Pharm. Pharmacol. 2015, 5, 411–415. [Google Scholar] [CrossRef]
  122. Sabogal-Guáqueta, A.M.; Muñoz-Manco, J.I.; Ramírez-Pineda, J.R.; Lamprea-Rodriguez, M.; Osorio, E.; Cardona-Gómez, G.P. The Flavonoid Quercetin Ameliorates Alzheimer’s Disease Pathology and Protects Cognitive and Emotional Function in Aged Triple Transgenic Alzheimer’s Disease Model Mice. Neuropharmacology 2015, 93, 134–145. [Google Scholar] [CrossRef] [PubMed]
  123. Tota, S.; Awasthi, H.; Kamat, P.K.; Nath, C.; Hanif, K. Protective Effect of Quercetin against Intracerebral Streptozotocin Induced Reduction in Cerebral Blood Flow and Impairment of Memory in Mice. Behav. Brain Res. 2010, 209, 73–79. [Google Scholar] [CrossRef]
  124. Choi, G.N.; Kim, J.H.; Kwak, J.H.; Jeong, C.-H.; Jeong, H.R.; Lee, U.; Heo, H.J. Effect of Quercetin on Learning and Memory Performance in ICR Mice under Neurotoxic Trimethyltin Exposure. Food Chem. 2012, 132, 1019–1024. [Google Scholar] [CrossRef]
  125. Wang, D.-M.; Li, S.-Q.; Wu, W.-L.; Zhu, X.-Y.; Wang, Y.; Yuan, H.-Y. Effects of Long-Term Treatment with Quercetin on Cognition and Mitochondrial Function in a Mouse Model of Alzheimer’s Disease. Neurochem. Res. 2014, 39, 1533–1543. [Google Scholar] [CrossRef] [PubMed]
  126. Beg, T.; Jyoti, S.; Naz, F.; Rahul; Ali, F.; Ali, S.K.; Reyad, A.M.; Siddique, Y.H. Protective Effect of Kaempferol on the Transgenic Drosophila Model of Alzheimer’s Disease. CNS Neurol. Disord.-Drug Targets-CNS Neurol. Disord. 2018, 17, 421–429. [Google Scholar] [CrossRef] [PubMed]
  127. Kouhestani, S.; Jafari, A.; Babaei, P. Kaempferol Attenuates Cognitive Deficit via Regulating Oxidative Stress and Neuroinflammation in an Ovariectomized Rat Model of Sporadic Dementia. Neural Regen. Res. 2018, 13, 1827–1832. [Google Scholar] [CrossRef] [PubMed]
  128. Babaei, P.; Eyvani, K.; Kouhestani, S. Sex-Independent Cognition Improvement in Response to Kaempferol in the Model of Sporadic Alzheimer’s Disease. Neurochem. Res. 2021, 46, 1480–1486. [Google Scholar] [CrossRef]
  129. Deshmukh, R.; Kaundal, M.; Bansal, V. Samardeep Caffeic Acid Attenuates Oxidative Stress, Learning and Memory Deficit in Intra-Cerebroventricular Streptozotocin Induced Experimental Dementia in Rats. Biomed. Pharmacother. 2016, 81, 56–62. [Google Scholar] [CrossRef] [PubMed]
  130. Khan, K.A.; Kumar, N.; Nayak, P.G.; Nampoothiri, M.; Shenoy, R.R.; Krishnadas, N.; Rao, C.M.; Mudgal, J. Impact of Caffeic Acid on Aluminium Chloride-Induced Dementia in Rats. J. Pharm. Pharmacol. 2013, 65, 1745–1752. [Google Scholar] [CrossRef]
  131. Chang, W.; Huang, D.; Lo, Y.M.; Tee, Q.; Kuo, P.; Wu, J.S.; Huang, W.; Shen, S. Protective Effect of Caffeic Acid against Alzheimer’s Disease Pathogenesis via Modulating Cerebral Insulin Signaling, β-Amyloid Accumulation, and Synaptic Plasticity in Hyperinsulinemic Rats. J. Agric. Food Chem. 2019, 67, 7684–7693. [Google Scholar] [CrossRef]
  132. Kadar, N.N.M.A.; Ahmad, F.; Teoh, S.L.; Yahaya, M.F. Comparable Benefits of Stingless Bee Honey and Caffeic Acid in Mitigating the Negative Effects of Metabolic Syndrome on the Brain. Antioxidants 2022, 11, 2154. [Google Scholar] [CrossRef]
  133. Wang, Y.; Wang, Y.; Li, J.; Hua, L.; Han, B.; Zhang, Y.; Yang, X.; Zeng, Z.; Bai, H.; Yin, H.; et al. Effects of Caffeic Acid on Learning Deficits in a Model of Alzheimer’s Disease. Int. J. Mol. Med. 2016, 38, 869–875. [Google Scholar] [CrossRef]
  134. Oboh, G.; Agunloye, O.M.; Akinyemi, A.J.; Ademiluyi, A.O.; Adefegha, S.A. Comparative Study on the Inhibitory Effect of Caffeic and Chlorogenic Acids on Key Enzymes Linked to Alzheimer’s Disease and Some Pro-Oxidant Induced Oxidative Stress in Rats’ Brain-In Vitro. Neurochem. Res. 2013, 38, 413–419. [Google Scholar] [CrossRef]
  135. Saenno, R.; Dornlakorn, O.; Anosri, T.; Kaewngam, S.; Sirichoat, A.; Aranarochana, A.; Pannangrong, W.; Wigmore, P.; Welbat, J.U. Caffeic Acid Alleviates Memory and Hippocampal Neurogenesis Deficits in Aging Rats Induced by D-Galactose. Nutrients 2022, 14, 2169. [Google Scholar] [CrossRef] [PubMed]
  136. Gao, L.; Li, X.; Meng, S.; Ma, T.; Wan, L.; Xu, S. Chlorogenic Acid Alleviates Aβ25-35-Induced Autophagy and Cognitive Impairment via the MTOR/TFEB Signaling Pathway. Drug Des. Devel. Ther. 2020, 14, 1705–1716. [Google Scholar] [CrossRef] [PubMed]
  137. Shen, W.; Qi, R.; Zhang, J.; Wang, Z.; Wang, H.; Hu, C.; Zhao, Y.; Bie, M.; Wang, Y.; Fu, Y.; et al. Chlorogenic Acid Inhibits LPS-Induced Microglial Activation and Improves Survival of Dopaminergic Neurons. Brain Res. Bull. 2012, 88, 487–494. [Google Scholar] [CrossRef] [PubMed]
  138. Kwon, S.-H.; Lee, H.-K.; Kim, J.-A.; Hong, S.-I.; Kim, H.-C.; Jo, T.-H.; Park, Y.-I.; Lee, C.-K.; Kim, Y.-B.; Lee, S.-Y.; et al. Neuroprotective Effects of Chlorogenic Acid on Scopolamine-Induced Amnesia via Anti-Acetylcholinesterase and Anti-Oxidative Activities in Mice. Eur. J. Pharmacol. 2010, 649, 210–217. [Google Scholar] [CrossRef] [PubMed]
  139. Kiasalari, Z.; Heydarifard, R.; Khalili, M.; Afshin-Majd, S.; Baluchnejadmojarad, T.; Zahedi, E.; Sanaierad, A.; Roghani, M. Ellagic Acid Ameliorates Learning and Memory Deficits in a Rat Model of Alzheimer’s Disease: An Exploration of Underlying Mechanisms. Psychopharmacology 2017, 234, 1841–1852. [Google Scholar] [CrossRef]
  140. Harakeh, S.; Qari, M.H.; Ramadan, W.S.; Al Jaouni, S.K.; Almuhayawi, M.S.; Al Amri, T.; Ashraf, G.M.; Bharali, D.J.; Mousa, S.A. A Novel Nanoformulation of Ellagic Acid Is Promising in Restoring Oxidative Homeostasis in Rat Brains with Alzheimer’s Disease. Curr. Drug Metab. 2021, 22, 299–307. [Google Scholar] [CrossRef]
  141. Zhong, L.; Liu, H.; Zhang, W.; Liu, X.; Jiang, B.; Fei, H.; Sun, Z. Ellagic Acid Ameliorates Learning and Memory Impairment in APP/PS1 Transgenic Mice via Inhibition of Β-amyloid Production and Tau Hyperphosphorylation. Exp. Ther. Med. 2018, 16, 4951–4958. [Google Scholar] [CrossRef]
  142. Mori, T.; Koyama, N.; Yokoo, T.; Segawa, T.; Maeda, M.; Sawmiller, D.; Tan, J.; Town, T. Gallic Acid Is a Dual α/β-Secretase Modulator That Reverses Cognitive Impairment and Remediates Pathology in Alzheimer Mice. J. Biol. Chem. 2020, 295, 16251–16266. [Google Scholar] [CrossRef]
  143. Ramadan, W.S.; Alkarim, S. Ellagic Acid Modulates the Amyloid Precursor Protein Gene via Superoxide Dismutase Regulation in the Entorhinal Cortex in an Experimental Alzheimer’s Model. Cells 2021, 10, 3511. [Google Scholar] [CrossRef]
  144. Jha, A.B.; Panchal, S.S.; Shah, A. Ellagic Acid: Insights into Its Neuroprotective and Cognitive Enhancement Effects in Sporadic Alzheimer’s Disease. Pharmacol. Biochem. Behav. 2018, 175, 33–46. [Google Scholar] [CrossRef]
  145. Wang, N.-Y.; Li, J.-N.; Liu, W.-L.; Huang, Q.; Li, W.-X.; Tan, Y.-H.; Liu, F.; Song, Z.-H.; Wang, M.-Y.; Xie, N.; et al. Ferulic Acid Ameliorates Alzheimer’s Disease-like Pathology and Repairs Cognitive Decline by Preventing Capillary Hypofunction in APP/PS1 Mice. Neurotherapeutics 2021, 18, 1064–1080. [Google Scholar] [CrossRef] [PubMed]
  146. Mori, T.; Koyama, N.; Tan, J.; Segawa, T.; Maeda, M.; Town, T. Combined Treatment with the Phenolics (−)-Epigallocatechin-3-Gallate and Ferulic Acid Improves Cognition and Reduces Alzheimer-like Pathology in Mice. J. Biol. Chem. 2019, 294, 2714–2731. [Google Scholar] [CrossRef]
  147. Yan, J.-J.; Jung, J.-S.; Kim, T.-K.; Hasan, M.A.; Hong, C.-W.; Nam, J.-S.; Song, D.-K. Protective Effects of Ferulic Acid in Amyloid Precursor Protein Plus Presenilin-1 Transgenic Mouse Model of Alzheimer Disease. Biol. Pharm. Bull. 2013, 36, 140–143. [Google Scholar] [CrossRef] [PubMed]
  148. Mori, T.; Koyama, N.; Guillot-Sestier, M.-V.; Tan, J.; Town, T. Ferulic Acid Is a Nutraceutical β-Secretase Modulator That Improves Behavioral Impairment and Alzheimer-like Pathology in Transgenic Mice. PLoS ONE 2013, 8, e55774. [Google Scholar] [CrossRef] [PubMed]
  149. Cho, J.-Y.; Kim, H.-S.; Kim, D.-H.; Yan, J.-J.; Lee, H.-K.; Suh, H.-W.; Song, D.-K. Inhibitory Effects of Long-Term Administration of Ferulic Acid on Astrocyte Activation Induced by Intracerebroventricular Injection of β-Amyloid Peptide (1–42) in Mice. Prog. Neuropsychopharmacol. Biol. Psychiatry 2005, 29, 901–907. [Google Scholar] [CrossRef]
  150. Kim, H.-S.; Cho, J.; Kim, D.-H.; Yan, J.-J.; Lee, H.-K.; Suh, H.-W.; Song, D.-K. Inhibitory Effects of Long-Term Administration of Ferulic Acid on Microglial Activation Induced by Intracerebroventricular Injection of β-Amyloid Peptide (1—42) in Mice. Biol. Pharm. Bull. 2004, 27, 120–121. [Google Scholar] [CrossRef] [PubMed]
  151. Kim, M.-J.; Seong, A.-R.; Yoo, J.-Y.; Jin, C.-H.; Lee, Y.-H.; Kim, Y.J.; Lee, J.; Jun, W.J.; Yoon, H.-G. Gallic Acid, a Histone Acetyltransferase Inhibitor, Suppresses β-Amyloid Neurotoxicity by Inhibiting Microglial-Mediated Neuroinflammation. Mol. Nutr. Food Res. 2011, 55, 1798–1808. [Google Scholar] [CrossRef]
  152. Ogunlade, B.; Adelakun, S.A.; Agie, J.A. Nutritional Supplementation of Gallic Acid Ameliorates Alzheimer-Type Hippocampal Neurodegeneration and Cognitive Impairment Induced by Aluminum Chloride Exposure in Adult Wistar Rats. Drug Chem. Toxicol. 2022, 45, 651–662. [Google Scholar] [CrossRef]
  153. Ogunsuyi, O.B.; Oboh, G.; Oluokun, O.O.; Ademiluyi, A.O.; Ogunruku, O.O. Gallic Acid Protects against Neurochemical Alterations in Transgenic Drosophila Model of Alzheimer’s Disease. Adv. Tradit. Med. 2020, 20, 89–98. [Google Scholar] [CrossRef]
  154. Hajipour, S.; Sarkaki, A.; Farbood, Y.; Eidi, A.; Mortazavi, P.; Valizadeh, Z. Effect of Gallic Acid on Dementia Type of Alzheimer Disease in Rats: Electrophysiological and Histological Studies. Basic Clin. Neurosci. 2016, 7, 97–106. [Google Scholar] [CrossRef]
  155. Mansouri, M.T.; Naghizadeh, B.; Ghorbanzadeh, B.; Farbood, Y.; Sarkaki, A.; Bavarsad, K. Gallic Acid Prevents Memory Deficits and Oxidative Stress Induced by Intracerebroventricular Injection of Streptozotocin in Rats. Pharmacol. Biochem. Behav. 2013, 111, 90–96. [Google Scholar] [CrossRef] [PubMed]
  156. Rashno, M.; Gholipour, P.; Salehi, I.; Komaki, A.; Rashidi, K.; Khoshnam, S.E.; Ghaderi, S. P-Coumaric Acid Mitigates Passive Avoidance Memory and Hippocampal Synaptic Plasticity Impairments in Aluminum Chloride-Induced Alzheimer’s Disease Rat Model. J. Funct. Foods 2022, 94, 105117. [Google Scholar] [CrossRef]
  157. Ghaderi, S.; Gholipour, P.; Komaki, A.; Salehi, I.; Rashidi, K.; Khoshnam, S.E.; Rashno, M. P-Coumaric Acid Ameliorates Cognitive and Non-Cognitive Disturbances in a Rat Model of Alzheimer’s Disease: The Role of Oxidative Stress and Inflammation. Int. Immunopharmacol. 2022, 112, 109295. [Google Scholar] [CrossRef] [PubMed]
  158. Kim, H.-B.; Lee, S.; Hwang, E.-S.; Maeng, S.; Park, J.-H. P-Coumaric Acid Enhances Long-Term Potentiation and Recovers Scopolamine-Induced Learning and Memory Impairments. Biochem. Biophys. Res. Commun. 2017, 492, 493–499. [Google Scholar] [CrossRef] [PubMed]
  159. Devi, S.; Kumar, V.; Singh, S.K.; Dubey, A.K.; Kim, J.-J. Flavonoids: Potential Candidates for the Treatment of Neurodegenerative Disorders. Biomedicines 2021, 9, 99. [Google Scholar] [CrossRef] [PubMed]
  160. Grosso, C.; Valentao, P.; Ferreres, F.; Andrade, P.B. The Use of Flavonoids in Central Nervous System Disorders. Curr. Med. Chem. 2013, 20, 4694–4719. [Google Scholar] [CrossRef]
  161. Caruso, G.; Godos, J.; Privitera, A.; Lanza, G.; Castellano, S.; Chillemi, A.; Bruni, O.; Ferri, R.; Caraci, F.; Grosso, G. Phenolic Acids and Prevention of Cognitive Decline: Polyphenols with a Neuroprotective Role in Cognitive Disorders and Alzheimer’s Disease. Nutrients 2022, 14, 819. [Google Scholar] [CrossRef]
  162. Freyssin, A.; Page, G.; Fauconneau, B.; Bilan, A.R. Natural Polyphenols Effects on Protein Aggregates in Alzheimer’s and Parkinson’s Prion-like Diseases. Neural Regen. Res. 2018, 13, 955. [Google Scholar] [CrossRef]
  163. Fernandes, M.Y.D.; Dobrachinski, F.; Silva, H.B.; Lopes, J.P.; Gonçalves, F.Q.; Soares, F.A.A.; Porciúncula, L.O.; Andrade, G.M.; Cunha, R.A.; Tomé, A.R. Neuromodulation and Neuroprotective Effects of Chlorogenic Acids in Excitatory Synapses of Mouse Hippocampal Slices. Sci. Rep. 2021, 11, 10488. [Google Scholar] [CrossRef]
  164. Matsui, T.; Ingelsson, M.; Fukumoto, H.; Ramasamy, K.; Kowa, H.; Frosch, M.P.; Irizarry, M.C.; Hyman, B.T. Expression of APP Pathway MRNAs and Proteins in Alzheimer’s Disease. Brain Res. 2007, 1161, 116–123. [Google Scholar] [CrossRef]
  165. Rezai-Zadeh, K.; Shytle, R.D.; Bai, Y.; Tian, J.; Hou, H.; Mori, T.; Zeng, J.; Obregon, D.; Town, T.; Tan, J. Flavonoid-Mediated Presenilin-1 Phosphorylation Reduces Alzheimer’s Disease β-Amyloid Production. J. Cell. Mol. Med. 2009, 13, 574–588. [Google Scholar] [CrossRef]
  166. Lubos, E.; Loscalzo, J.; Handy, D.E. Glutathione Peroxidase-1 in Health and Disease: From Molecular Mechanisms to Therapeutic Opportunities. Antioxid. Redox Signal. 2011, 15, 1957–1997. [Google Scholar] [CrossRef] [PubMed]
  167. Hole, K.L.; Williams, R.J. Flavonoids as an Intervention for Alzheimer’s Disease: Progress and Hurdles Towards Defining a Mechanism of Action. Brain Plast. 2020, 6, 167–192. [Google Scholar] [CrossRef] [PubMed]
  168. Jin, Y.; Fan, Y.; Yan, E.; Liu, Z.; Zong, Z.; Qi, Z. Effects of Sodium Ferulate on Amyloid-Beta-Induced MKK3/MKK6-P38 MAPK-Hsp27 Signal Pathway and Apoptosis in Rat Hippocampus. Acta Pharmacol. Sin. 2006, 27, 1309–1316. [Google Scholar] [CrossRef]
  169. Jang, S.; Kelley, K.W.; Johnson, R.W. Luteolin Reduces IL-6 Production in Microglia by Inhibiting JNK Phosphorylation and Activation of AP-1. Proc. Natl. Acad. Sci. USA 2008, 105, 7534–7539. [Google Scholar] [CrossRef]
  170. Zheng, Q.; Kebede, M.T.; Kemeh, M.M.; Islam, S.; Lee, B.; Bleck, S.D.; Wurfl, L.A.; Lazo, N.D. Inhibition of the Self-Assembly of Aβ and of Tau by Polyphenols: Mechanistic Studies. Molecules 2019, 24, 2316. [Google Scholar] [CrossRef] [PubMed]
  171. Hamaguchi, T.; Ono, K.; Murase, A.; Yamada, M. Phenolic Compounds Prevent Alzheimer’s Pathology through Different Effects on the Amyloid-β Aggregation Pathway. Am. J. Pathol. 2009, 175, 2557–2565. [Google Scholar] [CrossRef]
  172. Ochiai, R.; Saitou, K.; Suzukamo, C.; Osaki, N.; Asada, T. Effect of Chlorogenic Acids on Cognitive Function in Mild Cognitive Impairment: A Randomized Controlled Crossover Trial. J. Alzheimers Dis. 2019, 72, 1209–1216. [Google Scholar] [CrossRef]
  173. Kato, M.; Ochiai, R.; Kozuma, K.; Sato, H.; Katsuragi, Y. Effect of Chlorogenic Acid Intake on Cognitive Function in the Elderly: A Pilot Study. Evid. Based Complement. Alternat. Med. 2018, 2018, e8608497. [Google Scholar] [CrossRef]
  174. Saitou, K.; Ochiai, R.; Kozuma, K.; Sato, H.; Koikeda, T.; Osaki, N.; Katsuragi, Y. Effect of Chlorogenic Acids on Cognitive Function: A Randomized, Double-Blind, Placebo-Controlled Trial. Nutrients 2018, 10, 1337. [Google Scholar] [CrossRef]
  175. Root, M.; Ravine, E.; Harper, A. Flavonol Intake and Cognitive Decline in Middle-Aged Adults. J. Med. Food 2015, 18, 1327–1332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Azman, K.F.; Zakaria, R.; Aziz, C.B.A.; Othman, Z. Tualang Honey Attenuates Noise Stress-Induced Memory Deficits in Aged Rats. Oxid. Med. Cell. Longev. 2016, 2016, e1549158. [Google Scholar] [CrossRef] [PubMed]
  177. Azman, K.F.; Zakaria, R.; Othman, Z.; Aziz, C.B.A. Neuroprotective Effects of Tualang Honey against Oxidative Stress and Memory Decline in Young and Aged Rats Exposed to Noise Stress. J. Taibah Univ. Sci. 2018, 12, 273–284. [Google Scholar] [CrossRef]
  178. al-rahbi, B.; Zakaria, R.; Othman, Z.; Hassan, A.; Ahmad, A. The Effects of Tualang Honey Supplement on Medial Prefrontal Cortex Morphology and Cholinergic System in Stressed Ovariectomised Rats. Int. J. Appl. Res. Nat. Prod. 2014, 7, 28–36. [Google Scholar]
  179. Kamarulzaidi, M.A.; Yusoff, M.Y.Z.M.; Mohamed, A.M.; Adli, D.S.H. Tualang Honey Consumption Enhanced Hippocampal Pyramidal Count and Spatial Memory Performance of Adult Male Rats. Sains Malays. 2016, 45, 215–220. [Google Scholar]
  180. Al-Rahbi, B.; Zakaria, R.; Othman, Z.; Hassan, A.; Ismail, Z.I.M.; Muthuraju, S. Tualang Honey Supplement Improves Memory Performance and Hippocampal Morphology in Stressed Ovariectomized Rats. Acta Histochem. 2014, 116, 79–88. [Google Scholar] [CrossRef]
  181. Adeniyi, I.A.; Babalola, K.T.; Adekoya, V.A.; Oyebanjo, O.; Ajayi, A.M.; Onasanwo, S.A. Neuropharmacological Effects of Honey in Lipopolysaccharide-Induced Neuroinflammation, Cognitive Impairment, Anxiety and Motor Impairment. Nutr. Neurosci. 2022, 14, 1–14. [Google Scholar] [CrossRef]
  182. Cai, M.; Shin, B.Y.; Kim, D.H.; Kim, J.M.; Park, S.J.; Park, C.S.; Won, D.H.; Hong, N.D.; Kang, D.H.; Yutaka, Y.; et al. Neuroprotective Effects of a Traditional Herbal Prescription on Transient Cerebral Global Ischemia in Gerbils. J. Ethnopharmacol. 2011, 138, 723–730. [Google Scholar] [CrossRef]
  183. Oyefuga, O.H.; Ajani, E.O.; Salau, B.A.; Agboola, F.; Adebawo, O.O. Honey Consumption and Its Anti-Ageing Potency in White Wister Albino Rats. Sch. J. Biol. Sci. 2012, 1, 15–19. [Google Scholar]
  184. Djordjevic, J.; Jones-Gotman, M.; De Sousa, K.; Chertkow, H. Olfaction in Patients with Mild Cognitive Impairment and Alzheimer’s Disease. Neurobiol. Aging 2008, 29, 693–706. [Google Scholar] [CrossRef] [PubMed]
  185. Royet, J.P.; Croisile, B.; Williamson-Vasta, R.; Hibert, O.; Serclerat, D.; Guerin, J. Rating of Different Olfactory Judgements in Alzheimer’s Disease. Chem. Senses 2001, 26, 409–417. [Google Scholar] [CrossRef]
  186. Li, W.; Howard, J.D.; Gottfried, J.A. Disruption of Odour Quality Coding in Piriform Cortex Mediates Olfactory Deficits in Alzheimer’s Disease. Brain 2010, 133, 2714–2726. [Google Scholar] [CrossRef]
  187. Devanand, D.P.; Liu, X.; Tabert, M.H.; Pradhaban, G.; Cuasay, K.; Bell, K.; de Leon, M.J.; Doty, R.L.; Stern, Y.; Pelton, G.H. Combining Early Markers Strongly Predicts Conversion from Mild Cognitive Impairment to Alzheimer’s Disease. Biol. Psychiatry 2008, 64, 871–879. [Google Scholar] [CrossRef] [PubMed]
  188. Choudhury, N.; Chen, L.; Nguyen, V.T.; Al-Harthi, L.; Hu, X.-T. Medial Prefrontal Cortex Pyramidal Neurons Exhibit Functional Defects during Early Stage of Alzheimer’s Disease in 3xTg-AD Mice. Alzheimers Dement. 2021, 17, e057589. [Google Scholar] [CrossRef]
  189. Sun, Q.; Zhang, J.; Li, A.; Yao, M.; Liu, G.; Chen, S.; Luo, Y.; Wang, Z.; Gong, H.; Li, X.; et al. Acetylcholine Deficiency Disrupts Extratelencephalic Projection Neurons in the Prefrontal Cortex in a Mouse Model of Alzheimer’s Disease. Nat. Commun. 2022, 13, 998. [Google Scholar] [CrossRef]
  190. Liu, W.; Li, J.; Li, L.; Zhang, Y.; Yang, M.; Liang, S.; Li, L.; Dai, Y.; Chen, L.; Jia, W.; et al. Enhanced Medial Prefrontal Cortex and Hippocampal Activity Improves Memory Generalization in APP/PS1 Mice: A Multimodal Animal MRI Study. Front. Cell. Neurosci. 2022, 16, 848967. [Google Scholar] [CrossRef]
  191. Roostaei, T.; Nazeri, A.; Felsky, D.; De Jager, P.L.; Schneider, J.A.; Pollock, B.G.; Bennett, D.A.; Voineskos, A.N. Genome-Wide Interaction Study of Brain Beta-Amyloid Burden and Cognitive Impairment in Alzheimer’s Disease. Mol. Psychiatry 2017, 22, 287–295. [Google Scholar] [CrossRef]
  192. Basu, J.; Siegelbaum, S.A. The Corticohippocampal Circuit, Synaptic Plasticity, and Memory. Cold Spring Harb. Perspect. Biol. 2015, 7, a021733. [Google Scholar] [CrossRef] [PubMed]
  193. Buckley, M.J. The Role of the Perirhinal Cortex and Hippocampus in Learning, Memory, and Perception. Q. J. Exp. Psychol. Sect. B 2005, 58, 246–268. [Google Scholar] [CrossRef] [PubMed]
  194. Janeczek, M.; Gefen, T.; Samimi, M.; Kim, G.; Weintraub, S.; Bigio, E.; Rogalski, E.; Mesulam, M.-M.; Geula, C. Variations in Acetylcholinesterase Activity within Human Cortical Pyramidal Neurons Across Age and Cognitive Trajectories. Cereb. Cortex 2018, 28, 1329–1337. [Google Scholar] [CrossRef]
  195. Rogers, J.L.; Kesner, R.P. Cholinergic Modulation of the Hippocampus during Encoding and Retrieval. Neurobiol. Learn. Mem. 2003, 80, 332–342. [Google Scholar] [CrossRef] [PubMed]
  196. Jhamandas, J.H.; Cho, C.; Jassar, B.; Harris, K.; MacTavish, D.; Easaw, J. Cellular Mechanisms for Amyloid β-Protein Activation of Rat Cholinergic Basal Forebrain Neurons. J. Neurophysiol. 2001, 86, 1312–1320. [Google Scholar] [CrossRef] [PubMed]
  197. Kwakowsky, A.; Potapov, K.; Kim, S.; Peppercorn, K.; Tate, W.P.; Ábrahám, I.M. Treatment of Beta Amyloid 1–42 (Aβ1–42)-Induced Basal Forebrain Cholinergic Damage by a Non-Classical Estrogen Signaling Activator in Vivo. Sci. Rep. 2016, 6, 21101. [Google Scholar] [CrossRef] [PubMed]
  198. Zheng, W.-H.; Bastianetto, S.; Mennicken, F.; Ma, W.; Kar, S. Amyloid β Peptide Induces Tau Phosphorylation and Loss of Cholinergic Neurons in Rat Primary Septal Cultures. Neuroscience 2002, 115, 201–211. [Google Scholar] [CrossRef]
  199. Zambrzycka, A.; Alberghina, M.; Strosznajder, J.B. Effects of Aging and Amyloid-β Peptides on Choline Acetyltransferase Activity in Rat Brain. Neurochem. Res. 2002, 27, 277–281. [Google Scholar] [CrossRef]
  200. Vaucher, E.; Aumont, N.; Pearson, D.; Rowe, W.; Poirier, J.; Kar, S. Amyloid β Peptide Levels and Its Effects on Hippocampal Acetylcholine Release in Aged, Cognitively-Impaired and -Unimpaired Rats. J. Chem. Neuroanat. 2001, 21, 323–329. [Google Scholar] [CrossRef]
  201. Campanari, M.-L.; Navarrete, F.; Ginsberg, S.D.; Manzanares, J.; Sáez-Valero, J.; García-Ayllón, M.-S. Increased Expression of Readthrough Acetylcholinesterase Variants in the Brains of Alzheimer’s Disease Patients. J. Alzheimers Dis. 2016, 53, 831–841. [Google Scholar] [CrossRef]
  202. Berson, A.; Knobloch, M.; Hanan, M.; Diamant, S.; Sharoni, M.; Schuppli, D.; Geyer, B.C.; Ravid, R.; Mor, T.S.; Nitsch, R.M.; et al. Changes in Readthrough Acetylcholinesterase Expression Modulate Amyloid-Beta Pathology. Brain 2008, 131, 109–119. [Google Scholar] [CrossRef]
  203. Macdonald, I.R.; Maxwell, S.P.; Reid, G.A.; Cash, M.K.; DeBay, D.R.; Darvesh, S. Quantification of Butyrylcholinesterase Activity as a Sensitive and Specific Biomarker of Alzheimer’s Disease. J. Alzheimers Dis. 2017, 58, 491–505. [Google Scholar] [CrossRef]
  204. Li, Q.; Yang, H.; Chen, Y.; Sun, H. Recent Progress in the Identification of Selective Butyrylcholinesterase Inhibitors for Alzheimer’s Disease. Eur. J. Med. Chem. 2017, 132, 294–309. [Google Scholar] [CrossRef]
  205. Mushtaq, G.; Nigel, H.G.; Jalaluddin, A.K.; Mohammad, A.K. Status of Acetylcholinesterase and Butyrylcholinesterase in Alzheimer’s Disease and Type 2 Diabetes Mellitus. CNS Neurol. Disord.-Drug Targets-CNS Neurol. Disord. 2014, 13, 1432–1439. [Google Scholar] [CrossRef]
  206. Szwajgier, D.; Baranowska-Wójcik, E.; Winiarska-Mieczan, A.; Gajowniczek-Ałasa, D. Honeys as Possible Sources of Cholinesterase Inhibitors. Nutrients 2022, 14, 2969. [Google Scholar] [CrossRef] [PubMed]
  207. Greig, N.H.; Utsuki, T.; Ingram, D.K.; Wang, Y.; Pepeu, G.; Scali, C.; Yu, Q.-S.; Mamczarz, J.; Holloway, H.W.; Giordano, T.; et al. Selective Butyrylcholinesterase Inhibition Elevates Brain Acetylcholine, Augments Learning and Lowers Alzheimer β-Amyloid Peptide in Rodent. Proc. Natl. Acad. Sci. USA 2005, 102, 17213–17218. [Google Scholar] [CrossRef] [PubMed]
  208. Chepulis, L.M.; Starkey, N.J.; Waas, J.R.; Molan, P.C. The Effects of Long-Term Honey, Sucrose or Sugar-Free Diets on Memory and Anxiety in Rats. Physiol. Behav. 2009, 97, 359–368. [Google Scholar] [CrossRef] [PubMed]
  209. Azman, K.F.; Zakaria, R.; AbdAziz, C.; Othman, Z.; Al-Rahbi, B. Tualang Honey Improves Memory Performance and Decreases Depressive-like Behavior in Rats Exposed to Loud Noise Stress. Noise Health 2015, 17, 83–89. [Google Scholar] [CrossRef]
  210. Akouchekian, S.; Omranifard, V.; Maracy, M.R.; Pedram, A.; Zefreh, A.A. Efficacy of Herbal Combination of Sedge, Saffron, and Astragalus Honey on Major Neurocognitive Disorder. J. Res. Med. Sci. Off. J. Isfahan Univ. Med. Sci. 2018, 23, 58. [Google Scholar] [CrossRef]
  211. Mustafa, M.Z.; Zulkifli, F.N.; Fernandez, I.; Mariatulqabtiah, A.R.; Sangu, M.; Nor Azfa, J.; Mohamed, M.; Roslan, N. Stingless Bee Honey Improves Spatial Memory in Mice, Probably Associated with Brain-Derived Neurotrophic Factor (BDNF) and Inositol 1,4,5-Triphosphate Receptor Type 1 (Itpr1) Genes. Evid. Based Complement. Alternat. Med. 2019, 2019, e8258307. [Google Scholar] [CrossRef]
  212. Oyekunle, O.A.; Ogundeji, T.P.; Okojie, A.K. Behavioral Modifications Related to Consumption of a “Soft” Adaptogen, Bee Honey, by Rats. Neurophysiology 2011, 43, 38–41. [Google Scholar] [CrossRef]
  213. Warad, V.B.; Shastri, R.; Habbu, P.; Katti, P.; Jagannath, A.B.; Kulkarni, V.H.; Chakraborty, M. Preparation and Screening of Swarnaprashana for Nootropic Activity. Int. J. Nutr. Pharmacol. Neurol. Dis. 2014, 4, 170. [Google Scholar] [CrossRef]
  214. Akuchekian, S.; Layegh, E.; Najafi, M.; Barekatein, M.; Maracy, M.; Zomorodi, M.H. Effect of Herbal Medicine on Memory Impairment in Electroconvulsive Therapy (ECT)—ProQuest. J. Res. Med. Sci. 2012, 17, S59–S64. [Google Scholar]
  215. Badrasawi, M.M.; Shahar, S.; Manaf, Z.A.; Haron, H. Effect of Talbinah Food Consumption on Depressive Symptoms among Elderly Individuals in Long Term Care Facilities, Randomized Clinical Trial. Clin. Interv. Aging 2013, 8, 279–285. [Google Scholar] [CrossRef] [PubMed]
  216. Tsetsenis, T.; Badyna, J.K.; Wilson, J.A.; Zhang, X.; Krizman, E.N.; Subramaniyan, M.; Yang, K.; Thomas, S.A.; Dani, J.A. Midbrain Dopaminergic Innervation of the Hippocampus Is Sufficient to Modulate Formation of Aversive Memories. Proc. Natl. Acad. Sci. USA 2021, 118, e2111069118. [Google Scholar] [CrossRef] [PubMed]
  217. Rosen, Z.B.; Cheung, S.; Siegelbaum, S.A. Midbrain Dopamine Neurons Bidirectionally Regulate CA3-CA1 Synaptic Drive. Nat. Neurosci. 2015, 18, 1763–1771. [Google Scholar] [CrossRef] [PubMed]
  218. Kempadoo, K.A.; Mosharov, E.V.; Choi, S.J.; Sulzer, D.; Kandel, E.R. Dopamine Release from the Locus Coeruleus to the Dorsal Hippocampus Promotes Spatial Learning and Memory. Proc. Natl. Acad. Sci. USA 2016, 113, 14835–14840. [Google Scholar] [CrossRef]
  219. Smith, C.C.; Greene, R.W. CNS Dopamine Transmission Mediated by Noradrenergic Innervation. J. Neurosci. 2012, 32, 6072–6080. [Google Scholar] [CrossRef] [Green Version]
  220. Wagatsuma, A.; Okuyama, T.; Sun, C.; Smith, L.M.; Abe, K.; Tonegawa, S. Locus Coeruleus Input to Hippocampal CA3 Drives Single-Trial Learning of a Novel Context. Proc. Natl. Acad. Sci. USA 2018, 115, E310–E316. [Google Scholar] [CrossRef]
  221. Takeuchi, T.; Duszkiewicz, A.J.; Sonneborn, A.; Spooner, P.A.; Yamasaki, M.; Watanabe, M.; Smith, C.C.; Fernández, G.; Deisseroth, K.; Greene, R.W.; et al. Locus Coeruleus and Dopaminergic Consolidation of Everyday Memory. Nature 2016, 537, 357–362. [Google Scholar] [CrossRef]
  222. Duszkiewicz, A.J.; McNamara, C.G.; Takeuchi, T.; Genzel, L. Novelty and Dopaminergic Modulation of Memory Persistence: A Tale of Two Systems. Trends Neurosci. 2019, 42, 102–114. [Google Scholar] [CrossRef]
  223. Shohamy, D.; Wagner, A.D. Integrating Memories in the Human Brain: Hippocampal-Midbrain Encoding of Overlapping Events. Neuron 2008, 60, 378–389. [Google Scholar] [CrossRef]
  224. McNamara, C.G.; Tejero-Cantero, Á.; Trouche, S.; Campo-Urriza, N.; Dupret, D. Dopaminergic Neurons Promote Hippocampal Reactivation and Spatial Memory Persistence. Nat. Neurosci. 2014, 17, 1658–1660. [Google Scholar] [CrossRef]
  225. Lisman, J.E.; Grace, A.A. The Hippocampal-VTA Loop: Controlling the Entry of Information into Long-Term Memory. Neuron 2005, 46, 703–713. [Google Scholar] [CrossRef] [PubMed]
  226. Chowdhury, R.; Guitart-Masip, M.; Bunzeck, N.; Dolan, R.J.; Düzel, E. Dopamine Modulates Episodic Memory Persistence in Old Age. J. Neurosci. 2012, 32, 14193–14204. [Google Scholar] [CrossRef] [PubMed]
  227. Titulaer, J.; Björkholm, C.; Feltmann, K.; Malmlöf, T.; Mishra, D.; Gonzales, C.B.; Schilström, B.; Konradsson-Geuken, Å. The Importance of Ventral Hippocampal Dopamine and Norepinephrine in Recognition Memory. Front. Behav. Neurosci. 2021, 15, 667244. [Google Scholar] [CrossRef] [PubMed]
  228. Vargas, L.S.; Lima, K.R.; Ramborger, B.P.; Roehrs, R.; Izquierdo, I.; Mello-Carpes, P.B. Catecholaminergic Hippocampal Activation Is Necessary for Object Recognition Memory Persistence Induced by One-Single Physical Exercise Session. Behav. Brain Res. 2020, 379, 112356. [Google Scholar] [CrossRef]
  229. Peterson, A.C.; Zhang, S.; Hu, S.; Chao, H.H.; Li, C.R. The Effects of Age, from Young to Middle Adulthood, and Gender on Resting State Functional Connectivity of the Dopaminergic Midbrain. Front. Hum. Neurosci. 2017, 11, 52. [Google Scholar] [CrossRef] [Green Version]
  230. Noda, S.; Sato, S.; Fukuda, T.; Tada, N.; Hattori, N. Aging-Related Motor Function and Dopaminergic Neuronal Loss in C57BL/6 Mice. Mol. Brain 2020, 13, 46. [Google Scholar] [CrossRef]
  231. Simon, J.R.; Howard, J.H., Jr.; Howard, D.V. Adult Age Differences in Learning from Positive and Negative Probabilistic Feedback. Neuropsychology 2010, 24, 534–541. [Google Scholar] [CrossRef]
  232. Krashia, P.; Nobili, A.; D’Amelio, M. Unifying Hypothesis of Dopamine Neuron Loss in Neurodegenerative Diseases: Focusing on Alzheimer’s Disease. Front. Mol. Neurosci. 2019, 12, 123. [Google Scholar] [CrossRef]
  233. Nobili, A.; Latagliata, E.C.; Viscomi, M.T.; Cavallucci, V.; Cutuli, D.; Giacovazzo, G.; Krashia, P.; Rizzo, F.R.; Marino, R.; Federici, M.; et al. Dopamine Neuronal Loss Contributes to Memory and Reward Dysfunction in a Model of Alzheimer’s Disease. Nat. Commun. 2017, 8, 14727. [Google Scholar] [CrossRef]
  234. Cordella, A.; Krashia, P.; Nobili, A.; Pignataro, A.; La Barbera, L.; Viscomi, M.T.; Valzania, A.; Keller, F.; Ammassari-Teule, M.; Mercuri, N.B.; et al. Dopamine Loss Alters the Hippocampus-Nucleus Accumbens Synaptic Transmission in the Tg2576 Mouse Model of Alzheimer’s Disease. Neurobiol. Dis. 2018, 116, 142–154. [Google Scholar] [CrossRef]
  235. Moreno-Castilla, P.; Rodriguez-Duran, L.F.; Guzman-Ramos, K.; Barcenas-Femat, A.; Escobar, M.L.; Bermudez-Rattoni, F. Dopaminergic Neurotransmission Dysfunction Induced by Amyloid-β Transforms Cortical Long-Term Potentiation into Long-Term Depression and Produces Memory Impairment. Neurobiol. Aging 2016, 41, 187–199. [Google Scholar] [CrossRef] [PubMed]
  236. De Marco, M.; Venneri, A. Volume and Connectivity of the Ventral Tegmental Area Are Linked to Neurocognitive Signatures of Alzheimer’s Disease in Humans. J. Alzheimers Dis. 2018, 63, 167–180. [Google Scholar] [CrossRef] [PubMed]
  237. Huang, B.; Liu, J.; Ma, D.; Chen, G.; Wang, W.; Fu, S. Myricetin Prevents Dopaminergic Neurons from Undergoing Neuroinflammation-Mediated Degeneration in a Lipopolysaccharide-Induced Parkinson’s Disease Model. J. Funct. Foods 2018, 45, 452–461. [Google Scholar] [CrossRef]
  238. Bureau, G.; Longpré, F.; Martinoli, M.-G. Resveratrol and Quercetin, Two Natural Polyphenols, Reduce Apoptotic Neuronal Cell Death Induced by Neuroinflammation. J. Neurosci. Res. 2008, 86, 403–410. [Google Scholar] [CrossRef] [PubMed]
  239. Yang, S.; Zhang, D.; Yang, Z.; Hu, X.; Qian, S.; Liu, J.; Wilson, B.; Block, M.; Hong, J.-S. Curcumin Protects Dopaminergic Neuron Against LPS Induced Neurotoxicity in Primary Rat Neuron/Glia Culture. Neurochem. Res. 2008, 33, 2044–2053. [Google Scholar] [CrossRef] [PubMed]
  240. Bournival, J.; Plouffe, M.; Renaud, J.; Provencher, C.; Martinoli, M.-G. Quercetin and Sesamin Protect Dopaminergic Cells from MPP+-Induced Neuroinflammation in a Microglial (N9)-Neuronal (PC12) Coculture System. Oxid. Med. Cell. Longev. 2012, 2012, e921941. [Google Scholar] [CrossRef] [PubMed]
  241. Zhang, Y.; Wu, Q.; Zhang, L.; Wang, Q.; Yang, Z.; Liu, J.; Feng, L. Caffeic Acid Reduces A53T α-Synuclein by Activating JNK/Bcl-2-Mediated Autophagy in Vitro and Improves Behaviour and Protects Dopaminergic Neurons in a Mouse Model of Parkinson’s Disease. Pharmacol. Res. 2019, 150, 104538. [Google Scholar] [CrossRef]
  242. Kim, J.E.; Shrestha, A.C.; Kim, H.S.; Ham, H.N.; Kim, J.H.; Kim, Y.J.; Noh, Y.J.; Kim, S.J.; Kim, D.K.; Jo, H.K.; et al. WS-5 Extract of Curcuma Longa, Chaenomeles Sinensis, and Zingiber Officinale Contains Anti-AChE Compounds and Improves β-Amyloid-Induced Memory Impairment in Mice. Evid.-Based Complement. Altern. Med. ECAM 2019, 2019, 5160293. [Google Scholar] [CrossRef]
  243. He, X.; Yang, X.; Ou, R.; Ouyang, Y.; Wang, S.; Chen, Z.; Wen, S.; Pi, R. Synthesis and Evaluation of Multifunctional Ferulic and Caffeic Acid Dimers for Alzheimer’s Disease. Nat. Prod. Res. 2017, 31, 734–737. [Google Scholar] [CrossRef]
  244. Xu, S.; Sun, Y.; Dong, X. Design of Gallic Acid–Glutamine Conjugate and Chemical Implications for Its Potency Against Alzheimer’s Amyloid-β Fibrillogenesis. Bioconjug. Chem. 2022, 33, 677–690. [Google Scholar] [CrossRef]
  245. Jeon, S.G.; Song, E.J.; Lee, D.; Park, J.; Nam, Y.; Kim, J.; Moon, M. Traditional Oriental Medicines and Alzheimer’s Disease. Aging Dis. 2019, 10, 307–328. [Google Scholar] [CrossRef] [PubMed]
  246. Chen, M.; Du, Z.-Y.; Zheng, X.; Li, D.-L.; Zhou, R.-P.; Zhang, K. Use of Curcumin in Diagnosis, Prevention, and Treatment of Alzheimer’s Disease. Neural Regen. Res. 2018, 13, 742–752. [Google Scholar] [CrossRef] [PubMed]
  247. Mishra, S.; Palanivelu, K. The Effect of Curcumin (Turmeric) on Alzheimer’s Disease: An Overview. Ann. Indian Acad. Neurol. 2008, 11, 13–19. [Google Scholar] [CrossRef] [PubMed]
  248. Drummond, E.; Wisniewski, T. Alzheimer’s Disease: Experimental Models and Reality. Acta Neuropathol. (Berl.) 2017, 133, 155–175. [Google Scholar] [CrossRef] [PubMed]
  249. Kobayashi, H.; Murata, M.; Kawanishi, S.; Oikawa, S. Polyphenols with Anti-Amyloid β Aggregation Show Potential Risk of Toxicity Via Pro-Oxidant Properties. Int. J. Mol. Sci. 2020, 21, 3561. [Google Scholar] [CrossRef]
  250. Holland, T.M.; Agarwal, P.; Wang, Y.; Dhana, K.; Leurgans, S.E.; Shea, K.; Booth, S.L.; Rajan, K.; Schneider, J.A.; Barnes, L.L. Association of Dietary Intake of Flavonols With Changes in Global Cognition and Several Cognitive Abilities. Neurology 2022. [Google Scholar] [CrossRef]
  251. Shishtar, E.; Rogers, G.T.; Blumberg, J.B.; Au, R.; Jacques, P.F. Long-Term Dietary Flavonoid Intake and Risk of Alzheimer Disease and Related Dementias in the Framingham Offspring Cohort. Am. J. Clin. Nutr. 2020, 112, 343–353. [Google Scholar] [CrossRef]
  252. Agarwal, P.; Holland, T.M.; Wang, Y.; Bennett, D.A.; Morris, M.C. Association of Strawberries and Anthocyanidin Intake with Alzheimer’s Dementia Risk. Nutrients 2019, 11, 3060. [Google Scholar] [CrossRef]
  253. Devore, E.E.; Kang, J.H.; Breteler, M.M.B.; Grodstein, F. Dietary Intakes of Berries and Flavonoids in Relation to Cognitive Decline. Ann. Neurol. 2012, 72, 135–143. [Google Scholar] [CrossRef]
  254. Neveu, V.; Perez-Jiménez, J.; Vos, F.; Crespy, V.; du Chaffaut, L.; Mennen, L.; Knox, C.; Eisner, R.; Cruz, J.; Wishart, D.; et al. Phenol-Explorer: An Online Comprehensive Database on Polyphenol Contents in Foods. Database 2010, 2010, bap024. [Google Scholar] [CrossRef]
  255. Yahya, H.M.; Day, A.; Lawton, C.; Myrissa, K.; Croden, F.; Dye, L.; Williamson, G. Dietary Intake of 20 Polyphenol Subclasses in a Cohort of UK Women. Eur. J. Nutr. 2016, 55, 1839–1847. [Google Scholar] [CrossRef] [PubMed]
  256. García-Ayllón, M.-S.; Riba-Llena, I.; Serra-Basante, C.; Alom, J.; Boopathy, R.; Sáez-Valero, J. Altered Levels of Acetylcholinesterase in Alzheimer Plasma. PLOS ONE 2010, 5, e8701. [Google Scholar] [CrossRef] [PubMed]
  257. Nordberg, A.; Darreh-Shori, T.; Peskind, E.; Soininen, H.; Mousavi, M.; Eagle, G.; Lane, R. Different Cholinesterase Inhibitor Effects on CSF Cholinesterases in Alzheimer Patients. Curr. Alzheimer Res. 2009, 6, 4–14. [Google Scholar] [CrossRef] [PubMed]
  258. Davidsson, P.; Blennow, K.; Andreasen, N.; Eriksson, B.; Minthon, L.; Hesse, C. Differential Increase in Cerebrospinal Fluid-Acetylcholinesterase after Treatment with Acetylcholinesterase Inhibitors in Patients with Alzheimer’s Disease. Neurosci. Lett. 2001, 300, 157–160. [Google Scholar] [CrossRef]
  259. Guzmán-Ramos, K.; Moreno-Castilla, P.; Castro-Cruz, M.; McGaugh, J.L.; Martínez-Coria, H.; LaFerla, F.M.; Bermúdez-Rattoni, F. Restoration of Dopamine Release Deficits during Object Recognition Memory Acquisition Attenuates Cognitive Impairment in a Triple Transgenic Mice Model of Alzheimer’s Disease. Learn. Mem. 2012, 19, 453–460. [Google Scholar] [CrossRef] [Green Version]
  260. Tsunekawa, H.; Noda, Y.; Mouri, A.; Yoneda, F.; Nabeshima, T. Synergistic Effects of Selegiline and Donepezil on Cognitive Impairment Induced by Amyloid Beta (25–35). Behav. Brain Res. 2008, 190, 224–232. [Google Scholar] [CrossRef] [PubMed]
  261. Koch, G.; Di Lorenzo, F.; Bonnì, S.; Giacobbe, V.; Bozzali, M.; Caltagirone, C.; Martorana, A. Dopaminergic Modulation of Cortical Plasticity in Alzheimer’s Disease Patients. Neuropsychopharmacology 2014, 39, 2654–2661. [Google Scholar] [CrossRef]
  262. Nam, G.S.; Nam, K.-S.; Park, H.-J. Caffeic Acid Diminishes the Production and Release of Thrombogenic Molecules in Human Platelets. Biotechnol. Bioprocess Eng. 2018, 23, 641–648. [Google Scholar] [CrossRef]
  263. Abdel-Moneim, W.M.; Ghafeer, H.H. The Potential Protective Effect of Natural Honey Against Cadmium-Induced Hepatotoxicity and Nephrotoxicity. Mansoura J. Forensic Med. Clin. Toxicol. 2007, 15, 75–98. [Google Scholar] [CrossRef]
  264. Khalil, M.L.; Sulaiman, S.A. The Potential Role of Honey and Its Polyphenols in Preventing Heart Disease: A Review. Afr. J. Tradit. Complement. Altern. Med. 2010, 7, 315–321. [Google Scholar] [CrossRef]
  265. Eteraf-Oskouei, T.; Najafi, M. Traditional and Modern Uses of Natural Honey in Human Diseases: A Review. Iran. J. Basic Med. Sci. 2013, 16, 731–742. [Google Scholar] [PubMed]
Figure 1. The phenolic compounds, which comprises of flavonoids and phenolic acids.
Figure 1. The phenolic compounds, which comprises of flavonoids and phenolic acids.
Antioxidants 12 00427 g001
Figure 2. The possible effects of flavonoids in honey on the brain. Symbol (↑) represents increase while (↓) represents decrease.
Figure 2. The possible effects of flavonoids in honey on the brain. Symbol (↑) represents increase while (↓) represents decrease.
Antioxidants 12 00427 g002
Figure 3. The possible effects of phenolic acids in honey on the brain. Symbol (↑) represents increase while (↓) represents decrease.
Figure 3. The possible effects of phenolic acids in honey on the brain. Symbol (↑) represents increase while (↓) represents decrease.
Antioxidants 12 00427 g003
Table 1. Main flavonoid compounds affecting the physiological functioning and/or the pathophysiology of the central nervous system.
Table 1. Main flavonoid compounds affecting the physiological functioning and/or the pathophysiology of the central nervous system.
Flavonoid ComponentStudied ModelTesting MethodTime of Starting AdministrationPotential to Act AsStudied Region in BrainImportant Findings of the StudyReferences
MyricetinSTZ induced AD (Wistar) rat modelPassive avoidance test
IHC
1 day before stereotactic surgery (STZ exposure)Neuroprotective agentHippocampus (area CA3)Myricetin (at 10 mg/kg i.p.,) resulted in a better performance in avoidance test with decreased STL and increased TDC, along with increasing number of intact neurons in CA3 layer.[109]
Kunming MiceMWM test, and brain tissue analysisTogether with i.p. injection of scopolamineAntioxidant and anti-AChE agentHippocampusMyricetin decreased escape latency and increased time spent in target quadrant, and number of platform crossings.
Decreased the amount of MDA while improving antioxidant enzyme activities; it also sustained the concentration of ACh in the hippocampus.
[110]
Neurons from fetal rat cerebral cortex (E18)IHC, Immunoblotting, spectroscopy, and activity assays1 day before Aβ1–42 exposureAnti-amyloid and
neuroprotective agent
Not applicableMyrecetin protects neurons from Aβ1–42 induced injury and cell death.
It decreases production and aggregation of Aβ1–42 and Aβ1–40 (only at higher dose) that is also proved by increased activity of α-secretase and decreased activity of BACE1 (in a concentration-dependent manner).
[111]
LuteolinICV-STZ induced AD (Wistar) rat modelMWM task and probe tests;
IHC
3 days before injection of STZ Neuroprotective agentHippocampus (area CA1)Luteolin pre-treatment resulted in:
  • Decreased escape latency and travel distance to reach the hidden platform.
  • More time spent in the target quadrant.
  • More pyramidal cells in area CA1.
[112]
Sprague–Dawley rats (chronic hypoperfusion injury model)MWM task;
Brain tissue analysis
On 5th post-operative day of (bilateral common carotid arter) ligation surgeryAnti-inflammatory, antioxidant, and anti- amyloid agentCortex and hippocampusLuteolin-treated rats showed:
  • Decreased escape latency with more time spent in the target quadrant.
  • Decrease in the MDA level and an elevated SOD activity and the amount of GSH.
  • Decreased levels of TNF-α, IL-1β, and Aβ after luteolin treatment.
[113]
Adult male Balb/c mice;
Murine Neuro.2a, and LPS stimulated BV-2 (murine microglia cell line)
MWM task;
Brain tissue analysis
4 weeks before experimentAnti-inflammatory (aged mice), and neuroprotective (before LPS induction) agentHippocampusPretreatment with luteolin reduces pro-inflammatory mediators in microglia, therefore, prevents Neuro.2a cell death.
Reduction in mRNA levels of IL-1β and MHC class II, and TNFα (only with higher intake).
Better performance of aged mice fed with luteolin in MWM task.
[114]
Naringenin/NaringinHigh-fat-diet fed SAMP8 mice (a model of AD)MWM task and Barnes Maze test;
Brain tissue analysis
Along with the high-fat dietAnti-inflammatory, anti-amyloid, anti-tau, and neuroprotective agentCortex, hippocampus, and white matterNaringenin treatment resulted in:
  • Better performance in memory tasks.
  • Suppression of pro-inflammatory markers, and elevation of anti-inflammatory cytokines.
  • Reduction in the levels of soluble and insoluble Aβ40, Aβ42, APP and BACE1, p-tau, and GSK in hippocampus.
  • Reduced concentration of MDA, NO, and activity of SOD, GSH in cortex (compared to high-fat-diet fed group).
[115]
AlCl3+D-gal induced AD (Wistar) rat modelBehavioral tests;
Brain tissue analysis
Two weeks before AlCl3+D-gal inductionAntioxidant, anti-AChE, Serotonin- enhancer, and neuroprotective agentCortex and hippocampusNaringenin pre-treatment resulted in:
  • Better performance in memory tasks.
  • Decreased SOD activity and MDA levels, and increased activity of catalase, GPx, and GSH concentration.
  • Increased 5-HT and decreased 5-HIAA concentration.
  • Prevention of neuronal degeneration.
[116]
Intra- hippocampal Aβ1–40 induced (Wistar) rat modelY-maze, Radial arm maze task, passive avoidance test;
Brain tissue analysis
1 h before injecting Aβ1–40 bilaterally in the dorsal hippocampusAntioxidant and neuroprotective agentHippocampusPre-treatment of rats with naringenin caused:
  • Better performance in behavior tasks.
  • Lower level of MDA, without any significant difference in nitrit4e and SOD concentration.
  • Less DNA fragmentation (considered as a marker of apoptosis) in hippocampi.
[117]
ICV-STZ induced AD rat modelPassive avoidance test, MWM task;
Brain tissue analysis
14 days before ICV-STZ injectionAntioxidant and neuroprotective agentHippocampusPre-treatment with naringenin resulted in:
  • Better performance in behavioral tests.
  • Attenuation of lipid peroxidation and protein oxidation.
  • Increased level of GSH and increased activity of antioxidant enzymes.
  • Alleviation of Na+/K+-ATPase activity in hippocampus.
  • Restoration of ChAT neurons while maintaining normal morphology of the neurons in CA1.
[118]
Hydrocortisone injected AD mice modelMWM task NOR test and step-down test;$$$$$Brain tissue analysis21 days before hydrocortisone injectionAnti-amyloid, anti-tau, anti-AChE, antioxidant, and neuroprotective agentHippocampus and hypothalamusThe results of the pre-treatment of mice with Naringin were:
  • Better performance in behavioral tests.
  • The count, shape and distribution was similar to sham group.
  • Increased expression of estrogen receptor protein.
  • Decreased expression of p-Tau and CDK5 in hippocampus.
  • Inhibition of protein expression of Aβ, APP and BACE1 in hippocampus.
  • Increased ACh and ChAT, and decreased AChE in hippocampus.
  • Decreased levels of MDA and NO, and increased SOD in hippocampus.
[119]
QuercetinICR mice subjected to dexamethasoneMWM task3 h before dexamethasone i.p. injection Neuroprotective agentHippocampus (area CA3 and DG)More number of cells in DG in quercetin-treated group.[120]
ICV-STZ induced AD rat modelMWM taskAfter 1 week of ICV-STZ inductionNeuroprotective agentNot mentionedDecreased escape latency, and increased time spent in target quadrant.[121]
Homozygous 3xTg-AD
mice
MWM task, elevated plus maze;
Brain tissue analysis
Quercetin injected i.p., every 48 h for 3 months in AD mice before experimentationAnti-amyloid, anti-tau, anti-inflammatory and neuroprotective agentSubiculum, area CA1, entorrhinal cortex and amygdalaQuercertin-treated group showed:
  • An increased cell density in subiculum.
  • Decreased Aβ and tau fibrillary tangles deposition and in CA1, subiculum, and amygdala.
  • Significantly reduced astroglial and microglial immunoreactivity in the CA1 hippocampal area, the entorrhinal cortex, and the amygdala.
  • Improved memory in behavioral tests.
[122]
I.C.-STZ induced (Swiss) albino miceMWM task
Passive avoidance test;
Brain tissue analysis
Just after I.C.- STZ injectionAntioxidant and anti-AChE
agent
Whole brain (homogenate)Reduced mean latency in MWM task and increased TLT.
Reduction in MDA and nitrite levels, and inhibition of AChE activity (with higher dose of quercetin).
Increased GSH levels in quercetin-treated mice.
[123]
ICR mice subjected to TMT-induced neuronal deficitsY-maze and passive avoidance test;
Brain tissue analysis
21 days before the TMT inductionAntioxidant and
anti-AChE
agent
Whole brain (homogenate)Quercetin pre-treatment resulted in:
  • Improved performance in behavioral tests.
  • Inhibitory effect on AChE, with inhibition of lipid peroxidation (at a higher dose of Quercetin).
  • Antioxidant and radical scavenging ability shown by ABTS and FRAP assays.
[124]
APPswe/PS1dE9 (C57/BL) transgenic miceNOR test,
MWM test; Brain tissue analysis
16 weeks before sacrificeAntioxidant, anti-amyloid, and neuroprotective agentHippocampus and cortexMice treated with quercetin showed an increased recognition index in NOR test, decreased escape latency in MWM task.
Quercetin increases AMPK, prevents the formation of amyloid plaques, and alleviates hippocampal-mitochondria dysfunction.
[125]
KaempferolTransgenic Aβ flies (DS model)Climbing assay;
Brain tissue analysis
30 days before behavioral testsAntioxidant, anti-AChE, and neuroprotective agentWhole brain (homogenate)Dose-dependent increase in GSH content, and decrease in LPO, PC, GST, and AChE activity after kaempferol treatment compared with unexposed Aβ-flies.
Decreased apoptosis (evident by lower level of caspase enzymes) compared with the unexposed Aβ-flies.
[126]
Ovariectomized ICV-STZ induced AD (Wistar) rat modelMWM test;
Brain tissue analysis
On the same day as 2nd dose of STZ, and continued for 21 daysAntioxidant and anti-inflammatory agentHippocampusKaempferol consumption caused:
  • Reversal of STZ-induced cognitive dysfunction.
  • Enhanced hippocampal SOD and GSH levels.
  • Reduced levels of inflammatory markers MDA and TNF-α.
[127]
ICV-STZ induced AD (Wistar) rat model [128]
MWM = Morris Water Maze; NOR = novel object recognition; TLT = transfer latency time (i.e., the time taken to move from the open arm into any covered arm with four legs); ICR = Institute of Cancer Research; STZ = streptozotocin; ICV = intracerebroventricular; I.C. = intracerebral; IHC = immunohistochemistry; STL = step-through latency; TDL = time spent in the dark chamber; i.p. = intraperitoneal; TMT = trimethyltin; MDA = malondialdehyde; ABTS = 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid); 5-HIAA = 5-hydroxyindoleacetic acid; FRAP = ferric reducing antioxidant power; GSH = reduced glutatione; GPx = glutathione peroxidase; GST = glutathione-S-transferase; LPO = lipid peroxidation; PC = protein carbonyl content; ROS = reactive oxygen species; CDK5 = cyclin-dependent kinase 5; SAMP8 = senescence-accelerated mouse prone-8; NO = nitric oxide; ChAT = choline acetyltransferase; GSK3β = glycogen synthase kinase-3β; LPS = lipopolysaccharide; AMPK = AMP-activated protein kinase.
Table 2. Main phenolic acid compounds affecting the physiological functioning and/or the pathophysiology of the central nervous system.
Table 2. Main phenolic acid compounds affecting the physiological functioning and/or the pathophysiology of the central nervous system.
Phenolic Acid ComponentStudied ModelTesting MethodTime of Starting AdministrationPotential to Act AsStudied Region in BrainImportant Findings of StudyReferences
Caffeic acidICV-STZ induced AD (Wistar) rat modelMWM, NOR test and spontaneous locomotor activity;
Brain tissue analysis
1 h after first dose of ICV-STZAntioxidant and anti-AChE agentCerebral cortex and hippocampusCaffeic acid-treated rats showed:
  • Dose-dependent improvement in STZ-induced cognitive dysfunction.
  • Better memory (with a higher dose of caffeic acid), and more time spent in the target quadrant.
  • Dose-dependent attenuation of MDA, PC, and nitrite levels.
  • Restoration of depleted GSH and inhibition of AChE activity.
[129]
AlCl3-induced AD (Wistar) rat modelMWM;
Brain tissue analysis
20 days after AlCl3 (daily) injection Antioxidant and anti-AChE agentWhole brain (homogenate)Reversal of AlCl3 –induced memory deficits.
Inhibition of AChE activity, and nitrite levels in brain.
Increased levels of catalase, GSH, and GST in caffeic acid-treated group.
[130]
High-fat-diet-induced AD (Sprague-Dawley) rat modelMWM;
Brain tissue analysis
Along with the high-fat dietAntioxidant, anti-amyloid and$$$$$anti-tau agentCerebral cortex and hippocampusReversal of memory deficits in caffeic acid group.
Increased SOD, and decreased level of APP expression, β-Amyloid(1–42) content and BACE1 levels.
Decreased in p-Tau (Thr181) expression.
Increased synaptophysin expression in cortex, and drebrin expression after caffeic acid treatment.
[131]
High carbohydrate high fructose (HCHF) diet induced metabolic syndrome (Wistar) rat modelBrain tissue analysisAfter consumption of HCFC diet for 8 weeksAnti-inflammatory agentHippocampus (area CA1 and DG)Reduced TNF-α levels, and higher BDNF concentration compared with HCHF-only fed group.[132]
Intrahippocampally- Aβ1–40-induced AD (Sprague-Dawley) rat modelMWM;
Brain tissue analysis
After injecting Aβ1–40Antioxidant, anti-inflammatory, anti-AChE, and neuroprotective agentHippocampusCaffeic acid-treated group showed:
  • Decreased escape latency and mean path length, and more time spent in target quadrant.
  • Increased synaptophysin expression, and decreased AChE activity.
  • Decreased nitrite along with increased catalase and GSH levels.
  • Reduced NFκB-p65 expression with decreased activity of IL-6 and TNF-α.
  • Decreased p53 and P-p38 MAPK expression.
[133]
Wistar rats (whole brain in-vitro)Not applicableAdded to the supernatant of the homogenateAntioxidant,
anti-AChE, and anti-BChE agent
Whole brain homogenateAddition of caffeic acid caused:
  • Dose-dependant inhibition of AChE and BChE.
  • Dose-dependant decrease in MDA content.
  • High total antioxidant capacity and radical scavenging ability.
[134]
i.p. D-gal induced aging (Sprague-Dawley) rat modelNovel Object Location, NOR;
Brain tissue analysis
Along with D-galNeuroprotective agentHippocampusCo-treatment with caffeic acid displayed:
  • Dose-dependent attenuation of memory impairment.
  • Enhanced hippocampal neurogenesis by attenuation of reduced cell proliferation and increased survival of mature neurons.
[135]
Chlorogenic acidAPP/PS1 double transgenic miceMWM;Brain tissue analysisAt 3-month of ageNeuroprotective agentBrain (including histological evaluation of hippocampal CA1 area)Chlorogenic acid-treated mice showed:
  • Decrease in escape latency with more time spent in the target quadrant.
  • Cholorogenic acid protected against Aβ25–35 induced autophagy and promoted lysosomal function in APP/PS1 brain while restoring normal morphology of neurons in area CA1.
[136]
C57BL/6 mice
+
Primary neuro-glia cultures
Brain tissue analysis
+
Neuro-glia analysis and assays
7 days before LPS injection
+
2 h before incubation with LPS
Anti-inflammatory and
neuroprotective agent
Substantia nigraPre-treatment with chlorogenic acid:
  • Attenuated LPS-induced IL-1β and TNFα expression in Substantia nigra.
  • Inhibited nitrite and nitric oxide production along with attenuation of TNFα expression and NFκB signaling in LPS-stimulated microglia.
  • Protected dopaminergic neurons from microglia-mediated LPS toxicity.
[137]
Scopolamine-induced AD (ICR) mice modelY-maze test, passive avoidance test, MWM;
Brain tissue analysis
30-min before scopolamine injectionAntioxidant, and anti-AChE agentWhole brain (homogenate), and frontal cortex and hippocampus (homogenate)Cholorogenic acid pre-treatment resulted in:
  • Prevention of scopolamine-induced (short-term and long-term) learning and memory deficits.
  • Inhibition of AChE activity and MDA levels in hippocampus (at all tested doses) and frontal cortex(only at higher dose).
[138]
Ellagic acidIntrahippocampal microinjection Aβ25–35 induced
AD (Wistar) rat model
NOR, Y-maze, passive avoidance and radial arm maze tasks;
Brain tissue analysis
One week before Aβ-induction surgeryAnti-inflammatory, antioxidant, anti-AChE, and
neuroprotective agent
Hippocampus (including histological evaluation of CA1 area)Ellagic acid pre-treatment caused:
  • Improved discrimination ratio and memory performance.
  • Decreased MDA with an increase in GSH (at both doses), and catalase (only at higher dose).
  • Restored NFκB and nuclear/cytoplasmic ratio for Nrf2 (at both doses), and decreased TLR4 expression(only at higher dose).
  • Decreased level of AChE activity along with prevention of decline of CA1 neuronal count (at both doses).
[139]
AlCl3-induced AD (Wistar albino) rat modelNOR test;
Brain tissue analysis
After stopping AlCl3Antioxidant, anti-amyloid, anti-tau, and neuroprotective agentWhole brainThe results of ellagic acid treatment were:
  • Better memory in NOR test compared with untreated AD group.
  • Decreased lipid peroxidation along with increased levels of catalase, GSH, and total antioxidant capacity.
  • Improved neuronal morphology and lowering of amyloid and tau burden in cerebral cortex.
  • Co-treatment of ellagic acid with ellagic acid-loaded nanoparticles was more effective in mitigating all the behavioral and brain abnormalities.
[140]
APP/PS1 double-transgenic miceMWM;
Brain tissue analysis
1 week after acclimatizationAnti-amyloid, anti-tau, and neuroprotective agentHippocampusImproved learning and memory in the ellagic acid-treated group.
More number of neurons with reduced expression level of caspase-3 in hippocampus.
Decreased Aβ plaque deposition along with reduced levels of both Aβ40 and Aβ42 which is also confirmed by reduction in pThr668-APP and BACE1 expression.
Down-regulation of p-tau (pSer199-tau and pSer396-tau) by mediating AKT/GSK3β signaling pathway (increasing pSer473-AKT and lowering pTyr216-GSK3β).
[141]
NOR, Y-maze, radial arm water- maze tasks;
Brain tissue analysis
At 12 months of ageAntioxidant, anti-inflammatory, and anti-amyloid agent Whole brain (including study on EC, RSC, and hippocampus)Treatment with ellagic acid resulted in:
  • Complete reversal of learning and memory impairments and (anxiety-like) behavioral abnormalities.
  • Upregulation of α-secretase and downregulation of BACE1 with reduced Aβ-plaque deposition (both Aβ1–40 and Aβ1–42), and CAA.
  • Decreased number of immunoreactive glia (astroglia and microglia) with reduced expression of SOD1 and GPx1.
[142]
Oral AlCl3- induced AD (Wistar) rat modelNOR test;
Brain tissue analysis
4 weeks after the beginning of oral AlCl3 dosageAntioxidant, anti-amyloid, anti-tau, and neuroprotective agentECEllagic acid-treated group showed:
  • Improved discrimination index in NOR test.
  • Increased serum SOD (due to upregulated gene expression), GSH levels and higher mean total antioxidant capacity with decreased levels of TBRS (products of lipid peroxidation).
  • Restored thickness of EC with more neurons having normal morphology.
  • Down-regulation of APP and caspase-3 expression with reduced load of NFTs.
[143]
ICV-STZ induced AD (Wistar) rat modelRadial arm maze and Y-maze tasks;
Brain tissue analysis
1 day after STZ administrationAntioxidant, anti-inflammatory, anti-amyloid, and neuroprotective agentCerebral cortex (homogenate), EC and hippocampus proper (area CA1, CA2, CA3, and DG)Ellagic acid treatment caused:
  • Improved memory and cognitive scores.
  • Reduced levels of MDA and CRP together with elevated GSH and catalase activity.
  • Neurons having normal morphology with decreased Aβ-plaque burden in EC and hippocampus proper.
  • Reduction of immunoreactive astroglia and elevation of synaptophysin levels.
[144]
Ferulic acid APP/PS1 (transgenic) mice MWM task;
Brain tissue analysis
In AD mice of 6 months ageAnti-amyloid,
neurovascular protective agent
Whole brain (including study on cerebral cortex and hippocampus) Ferulic acid treatment effects on APP/PS1 mice were:
  • Restoration of learning and memory impairment.
  • Increased density of whole-brain blood vessels (including hippocampus) with prevention of reduction of diameter of hippocampal capillaries and, therefore, cerebral blood flow.
  • Reduction of Aβ plaque deposition in hippocampus(both Aβ1–42 and Aβ1–40) and cortex along with attenuated BACE1 activity.
  • Reduced microglia aggregates surrounding Aβ plaques.
[145]
NOR, Y-maze, radial arm-water maze tasks;
Brain tissue analysis
In 1-year-old AD- mice modelAnti-inflammatory, antioxidant, and anti-amyloid agentWhole brain (including study on RSC, EC, hippocampus)Treatment with Ferulic acid resulted in:
  • Reduction in cerebral amyloidosis and CAA.
  • Decreased reactive gliosis with reduced expression of SOD1, GPx1, TNF-α and IL-1β.
  • Elevated synaptophysin immunoreactivity in area CA1 and EC.
  • Moreover, the combination therapy of Ferulic acid with epigallocatechin-3-gallate was more effective and completely reversed all the behavioral and brain abnormalities.
[146]
NOR and Y-maze tasks;
Brain tissue analysis
In 6-month old AD miceAnti-amyloid, anti-inflammatory agentFrontal cortex and hippocampus Ferulic acid treatment:
  • Improved memory performance in NOR after low-dose (5.3 mg/kg/day) treatment, whereas treatment at a higher dose (16 mg/kg/day) was ineffective.
  • Reduction in cortical Aβ1–40 and Aβ1–42 levels (more effective at lower dose) with alleviation of IL-1β levels (at both doses).
[147]
PSAPP mice (AD-model)NOR, Y-maze and MWM tasks;
Brain tissue analysis
In 6-month-old AD miceAntioxidant, anti-inflammatory and anti-amyloid agentCingulate cortex, EC and hippocampus
+
Whole brain (homogenate)
Ferulic acid-treated PSAPP mice displayed:
  • Remediation of learning, memory and behavior impairment.
  • Reduction of cerebral amyloid burden and CAA by attenuating both Aβ1–40 and Aβ1–42 levels along with decreased BACE1 activity (at the translational/protein level).
  • Attenuation of glial activation with reducing expression of TNF-α, IL-1β, SOD1, catalase and GPx1.
[148]
ICR mice (ICV-induced Aβ1–42 AD-model)Brain tissue analysis4 weeks before ICV injection of Aβ1–42Antioxidant and anti-inflammatory agentHippocampusFerulic acid pre-treatment mitigated oxidative stress and neuroinflammation by blocking astroglial activation evident by double staining of 3-nitrotyrosine and endothelial nitric oxide synthase immunoreactive cells with GFAP.[149]
Brain tissue analysis4 weeks before ICV injection of Aβ1–42Antioxidant and anti-inflammatory agentHippocampusFerulic acid pre-treatment inhibited microglial activation evident by blocking of OX-42 (marker of activated microglia) and IFN-γ immunoreactivity.[150]
Gallic acidICV Aβ1–42 induced AD (ICR) mice modelY-maze and passive avoidance test;
Brain tissue analysis
3-weeks before ICV injection of AβAnti-inflammatory and neuroprotective agentWhole brain (homogenate), Cerebral cortex and hippocampusGallic acid pre-treatment:
  • Prevented Aβ-induced cognitive deficits.
  • Restored cytokine (iNOS and COX-2) levels in the cerebral cortex and hippocampus induced with Aβ and decreased neuronal apoptosis.
  • Inhibited nuclear translocation and acetylation of NFκB and prevented subsequent IL-1β release in mice brain.
[151]
Oral AlCl3-induced AD (Wistar) rat model Y-maze and MWM tests;
Brain tissue analysis
Together with AlCl3Antioxidant and neuroprotective agentHippocampusGallic acid co-ingestion group showed:
  • Improved learning and memory indices in behavioral tests.
  • Elevated levels of catalase, SOD and GSH with more neurons in hippocampus.
[152]
APP: BACE [high] transgenic Drosophila AD-modelBrain homogenate analysis5 days before sacrificeAntioxidant,
anti-inflammatory,
anti-AChE, and anti-BChE, and anti-amyloid agent
Whole brain (homogenate)Gallic acid caused:
  • Decreased BACE1 and Cholinesterase (AChE and BChE) activity.
  • Reduced MDA levels and increased catalase activity together with amelioration of reactive oxygen species burden.
  • Raised total thiol content.
[153]
Intrahippocampal Aβ1–42 induced
AD (Wistar) rat model
Electrophysiological analysis;
Brain tissue analysis
The 2nd day after intrahippocampal injectionAnti-amyloidHippocampus (area CA1 and DG)Improved amplitude and area under curve of LTP as recorded from DG in gallic acid-treated flies.
Reduced burden of Aβ plaques in area CA1 in treated group.
[154]
ICV-STZ
induced AD (Wistar) rat model
Passive avoidance and MWM tests;
Brain tissue analysis
5 days before ICV-STZ injectionAntioxidant agentCerebral cortex and hippocampusPre-treatment with Gallic acid resulted in:
  • Improved learning and memory in behavioral tests.
  • Decreased levels of MDA and increased total thiol levels with restoration of SOD, GPx and catalase activity.
[155]
p-Coumaric acid (p-CA)AlCl3-induced AD (Wistar) rat modelPassive avoidance test;
Electrophysiological analysis;
Histological analysis
1-h prior to AlCl3-induction Anti-amyloid and neuroprotective agentHippocampusp-CA pretreatment resulted in:Improved memory retrieval in behavioral test.
Mitigation of LTP impairment that is evident by increased amplitude and area under curve for the population spike, and the field excitatory postsynaptic potentials slope in electrophysiological recordings.
Reduced burden of Aβ-plaques in DG.
[156]
OFT, elevated plus maze, MWM, and forced swimming tests;
Brain tissue analysis
Antioxidant, anti-inflammatory, and neuroprotective agentCerebral cortex and hippocampus (histological studies on area CA1, CA3 and DG)p-CA pre-treatment improved memory, decreased anxiety and depression-like behavior, and increased exploratory activity.
p-CA increased SOD, GPx, and Catalase activities and decreased MDA NFκB, TNF-α, IL-1β, and IL-6 levels.
Increased number of intact neurons.
[157]
Scopolamine-induced AD (Sprague-Dawley) rat modelPassive avoidance and MWM test
Electrophysiological analysis
1-h before scopolamine administrationNeuroprotective agentHippocampus (area CA1)p-CA causes the following changes:
  • Increases amplitude of LTP with increased field excitatory postsynaptic potentials in electrophysiological studies in p-CA group.
  • Inhibition of NMDA receptor and AMPA receptor blockade and resultant increase in LTP.
  • Inhibits muscarinic receptor blockade which is proposed as the likely reason behind memory improvement in behavioral test.
[158]
EC = entorhinal cortex; RSC = retrosplenal cortex; DG = dentate gyrus; CC = cerebral cortex; CA = cornu ammonis; NFTs = neurofibrillary tangles; CAA = cerebral amyloid angiopathy; LTP = long-term potentiation; NOR = novel object recognition; OFT = open field test; LPS = lipopolysaccharide; TLR4 = Toll-like receptor 4, Nrf2 = nuclear factor (erythroid-derived) 2; NFκB = nuclear factor-kappa B; SOD = superoxide dismutase; PC = protein carbonyl content; BACE = β-Site APP-cleaving enzyme; GPx = glutathione peroxidase; TBRS = thiobarbituric-acid-reactive substances; AlCl3 = aluminum chloride; IFN-γ-gamma interferon; NMDA receptor = N-methyl-D-aspartate receptor; AMPA receptor = α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor.5. Effect of honey as a neuroprotective agent.
Table 3. Effects of honey intake on memory and cognition in animal model.
Table 3. Effects of honey intake on memory and cognition in animal model.
Studied ModelDosageDuration of ExposureFindingsReference
Wistar rats induced with metabolic syndrome effects (MetS) by feeding high carbohydrate high fructose (HCHF) diet 15 mL of Kelulut honey dissolved in 15 mL of distilled water given at 0.1 mL/kg of body weight daily35 days (after 16 weeks of HCFC diet) Reduced anxiety compared with the MetS group
Enhanced memory efficiency than both control and MetS groups
Increased number of pyramidal cells in the hippocampus compared with the MetS group
[34]
Female Swiss albino mice
(age = 2.5 months)
A total concentration of 750 mg/kg and 2000 mg/kg among groups—with 0.3 mL of Stingless bee honey dilution forced-fed daily7 days (acute) and 35 days (semichronic)Improvement in learning and spatial reference memory[211]
Stressed ovariectomized Sprague–Dawley rats
(approximately 8 weeks old)
0.2 g/kg body weight daily of Tualang honey diluted with 1ml of distilled water18 days (started 3 days before stress induction)Decreased anxiety-like behavior in the stressed ovariectomized rats (StOE) that is comparable to β-estradiol (E2) treatment[29]
Improved short-term and long-term memory in the StOE rats
Increased pyramidal cells in CA2, CA3, and DG of hippocampus in the StOE rats
Results were comparable to that after E2 treatment
[180]
Sprague–Dawley rats
(approximately 2 months old)
Experimental diet of containing 100 g/kg honeydew honey that was available ad libitum3, 6, 9, and 12 monthsImproved spatial memory and decreased anxiety-like behavior[208]
Sprague–Dawley rats
(2 months old)
0.2 g/kg body weight Tualang honey dissolved in distilled water/daily35 daysImproved short-term and long-term memory
Decreased depressive-like symptoms
[209]
Sprague–Dawley rats
(16 months old)
200 mg/kg body weight of Tualang honey/daily28 days (started 14 days prior to stress procedures)Improvement in both short-term and long-term memory
More Nissl-positive cells in the mPFC and hippocampus
Greater number of pyramidal cells in the mPFC and hippocampus exhibiting normal shape and structure
[176]
Male Sprague–Dawley rats
Young (2 months old), and aged (16 months old)
200 mg/kg body weight of Tualang honey/daily28 days (started 14 days prior to stress procedures)[177]
Wistar rats0.5, 1.0, and 2.0 g/kg body weightSingly dose (1 h before behavioral tests)Honey, in a dose-dependent manner, ameliorates the anxiety-like behavior and possibly also acts as an anti-depressant[212]
Swiss albino mice
Young (3–4 months), aged (12–15 months)
The formulation containing honey (400 mL), ghee (800 mL) and gold (288 mg) was given at the dose of 30 mg/kg/daily15 daysThe formulation intake improved learning and memory in young and aged mice
Decreased activity of AChE in brain
[213]
Table 4. Effects of honey consumption on memory and cognition in humans.
Table 4. Effects of honey consumption on memory and cognition in humans.
SubjectsType of StudyDosageDuration of ExposureFindingsReference
Mild cognitively impaired, and
Cognitively intact controls
(all over 65 years old)
Randomized, placebo control, double-blind1 tablespoon daily5 years—tested every 6 monthsAbout 28% of the placebo-given subjects, while less than 7% of the honey-ingested subjects developed dementia.[37]
Postmenopausal women
(aged between 45 and 60 years)
Cohort20 g Tualang honey (sachet) daily16 weeksImproved learning and memory scores in the auditory verbal learning test[38]
Patients diagnosed with mood disorders and candidates for electroconvulsive therapy (ECT)
(aged > 18 years)
Randomized, double-blind9 g of herbal combination of Crocus sativus, Cyperus rotundus, and honey/twice daily40 days (after initiation of ECT)Improvement of ECT-induced memory improvement, especially after one to two months of the last ECT session.[214]
Depressed elderly
individuals (aged 60 or more)
Crossover randomized 25 g Talbinah honey in 100 mL of water daily6 weeks (3 weeks + 1 week break + 3 weeks)Improvement in depression, stress, and mood disturbances scores.[215]
Patients diagnosed with mild to moderate major neurocognitive disorderRandomized, double-blind10 g Asparagalus honey with 1000 mg of sedge and 60 mg of saffron extracts daily3 monthsImproved attention, memory and cognition compared with the placebo-given group.[210]
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

Shaikh, A.; Ahmad, F.; Teoh, S.L.; Kumar, J.; Yahaya, M.F. Honey and Alzheimer’s Disease—Current Understanding and Future Prospects. Antioxidants 2023, 12, 427. https://doi.org/10.3390/antiox12020427

AMA Style

Shaikh A, Ahmad F, Teoh SL, Kumar J, Yahaya MF. Honey and Alzheimer’s Disease—Current Understanding and Future Prospects. Antioxidants. 2023; 12(2):427. https://doi.org/10.3390/antiox12020427

Chicago/Turabian Style

Shaikh, Ammara, Fairus Ahmad, Seong Lin Teoh, Jaya Kumar, and Mohamad Fairuz Yahaya. 2023. "Honey and Alzheimer’s Disease—Current Understanding and Future Prospects" Antioxidants 12, no. 2: 427. https://doi.org/10.3390/antiox12020427

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