Next Article in Journal
Quercetin Antagonizes the Sedative Effects of Linalool, Possibly through the GABAergic Interaction Pathway
Next Article in Special Issue
Bilobalide Induces Apoptosis in 3T3-L1 Mature Adipocytes through ROS-Mediated Mitochondria Pathway
Previous Article in Journal
NiCo-MOF Nanospheres Created by the Ultra-Fast Microwave Method for Use in High-Performance Supercapacitors
Previous Article in Special Issue
A Rational Combination of Cyclocarya paliurus Triterpene Acid Complex (TAC) and Se-Methylselenocysteine (MSC) Improves Glucose and Lipid Metabolism via the PI3K/Akt/GSK3β Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tracing the Path between Mushrooms and Alzheimer’s Disease—A Literature Review

1
Ciências Químicas e das Biomoléculas, Escola Superior de Saúde, Instituto Politécnico do Porto, 4200-072 Porto, Portugal
2
CIIMAR/CIMAR—Centro Interdisciplinar de Investigação Marinha e Ambiental, Universidade do Porto, 4450-208 Matosinhos, Portugal
3
REQUIMTE/LAQV, Instituto Superior de Engenharia do Porto, Instituto Politécnico do Porto, 4249-015 Porto, Portugal
4
TBIO—Centro de Investigação em Saúde Translacional e Biotecnologia Médica, Escola Superior de Saúde, Instituto Politécnico do Porto, 4200-072 Porto, Portugal
5
Departamento de Biologia, Faculdade de Ciências, Universidade do Porto, 4169-007 Porto, Portugal
6
CIMO, Centro de Investigação de Montanha, Instituto Politécnico de Bragança, 5300-253 Bragança, Portugal
7
Laboratório Associado para a Sustentabilidade e Tecnologia em Regiões de Montanha (SusTEC), Instituto Politécnico de Bragança, 5300-253 Bragança, Portugal
8
CISA, Centro de Investigação em Saúde e Ambiente, Escola Superior de Saúde, Instituto Politécnico do Porto, 4200-072 Porto, Portugal
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(14), 5614; https://doi.org/10.3390/molecules28145614
Submission received: 3 July 2023 / Revised: 19 July 2023 / Accepted: 21 July 2023 / Published: 24 July 2023
(This article belongs to the Special Issue Natural Compounds for Disease and Health)

Abstract

:
Alzheimer’s disease (AD) is well-known among neurodegenerative diseases for the decline of cognitive functions, making overall daily tasks difficult or impossible. The disease prevails as the most common form of dementia and remains without a well-defined etiology. Being considered a disease of multifactorial origin, current targeted treatments have only managed to reduce or control symptoms, and to date, only two drugs are close to being able to halt its progression. For decades, natural compounds produced by living organisms have been at the forefront of research for new therapies. Mushrooms, which are well-known for their nutritional and medicinal properties, have also been studied for their potential use in the treatment of AD. Natural products derived from mushrooms have shown to be beneficial in several AD-related mechanisms, including the inhibition of acetylcholinesterase (AChE) and β-secretase (BACE 1); the prevention of amyloid beta (Aβ) aggregation and neurotoxicity; and the prevention of Tau expression and aggregation, as well as antioxidant and anti-inflammatory potential. Several studies in the literature relate mushrooms to neurodegenerative diseases. However, to the best of our knowledge, there is no publication that summarizes only AD data. In this context, this review aims to link the therapeutic potential of mushrooms to AD by compiling the anti-AD potential of different mushroom extracts or isolated compounds, targeting known AD-related mechanisms.

Graphical Abstract

1. Introduction

Neurodegeneration comprises a series of illnesses that cause the gradual impairment of cognitive and motor abilities. Alzheimer’s disease (AD) is a neurodegenerative disease and the most common form of dementia (60–80% of worldwide dementia cases) [1]. Statistics show that in 2018 more than 8 million people over the age of 30 suffered from AD in Europe (the number is expected to double by 2050), and in the USA, data from 2023 point to nearly 6.5 million persons over 65 years to be affected with AD (incidence of 1 in more than 8 million people over the age of 30 suffered from AD in Europe (the number is expected to double by 2050), and in the USA, data from 2023 point to nearly 6.5 million persons over 65 years to be affected with AD (incidence of 1 in 9) [2,3] Linked to the incidence rate and lack of effective treatment, the high degree of disability and dependency of patients burden personal, medical, and social systems with excessive social and economic costs, which are expected to reach USD 2 trillion by 2030 on a global basis and EUR 330 billion by 2050 in Europe alone [4].
AD affects an individual’s ability to perform normal, daily activities, causing dependence, disability, and ultimately, death. It is thought that the disease can be present several years before manifesting any symptoms and, consequently, AD patients will only experience symptoms after a certain level of modification of the brain structure and activity [5]. At a macroscopic level, structural modifications include the age-dependent atrophy of the hippocampus and cerebral cortex, due to neuron degeneration. At a microscopic level, AD hallmarks include the formation of beta-amyloid (Aβ) plaques, due to the aggregation and deposition of Aβ peptide, and neurofibrillary tangles generated by the accumulation of the hyperphosphorylated Tau protein [6,7].
Despite being initially documented more than a century ago, AD lacks an established theory explaining its etiology, progression, and pathogenesis, as well as a treatment that is able to effectively halt disease progression or treat the cause. Nevertheless, some hypotheses have been proposed to explain the origin of AD, namely, the amyloid cascade hypothesis, the cholinergic hypothesis, the tau hypothesis, and the glutaminergic hypothesis [8]. In the first of these hypotheses, the Aβ peptide is produced in excess as a result of the β-secretase (BACE1) cleavage of the Aβ precursor protein (APP). Excessive Aβ undergoes a series of self-assembly steps, resulting in insoluble Aβ aggregates that are deposited extracellularly near the synapses. According to this hypothesis, Aβ aggregates trigger a cascade of events, leading to synaptic dysfunction and neuronal death, manifested in memory loss, and impaired cognitive functions [9]. In the cholinergic hypothesis, the involvement of the acetylcholinesterase enzyme (AChE) is a key factor. Cholinergic neurons in the central nervous system are involved in cognitive processes, such as memory and learning. The stimulation of receptors is caused by acetylcholine (ACh). One of the potential therapeutic strategies to AD is to increase the ACh levels in the synaptic clefts by inhibiting the biological activity of AChE [9,10]. The tau hypothesis relies on the hyperphosphorylation of the microtubule-associated protein tau. When hyperphosphorylated, tau is prone to aggregation and is responsible for neurofibrillary tangle formation. Its intracellular accumulation is associated with reduced neuronal function, impaired synaptic function, and neurotoxicity [6,11]. The glutaminergic hypothesis suggests the inappropriate stimulation of the N-methyl-D-aspartate receptors (NMDARs) as causing the deterioration of synaptic function. NMDARs are involved in synaptic transmission and plasticity. The hyperexcitation of these receptors causes excitotoxicity, which is mediated by increased Ca2+, resulting in synaptic dysfunction and neuronal death [12]. Moreover, epidemiological and translational research suggest that AD might also be promoted by neuroinflammation, systemic inflammation, and oxidative stress. Chronic inflammation has been observed as an early event in AD development [13]. In this case, neuroinflammation is induced through the accumulation of Aβ plaques that activate and recruit both microglia and reactive astrocytes. Glial cells are constantly being activated by the excess of Aβ plaques, resulting in the release of inflammatory cytokines [14]. Oxidative stress has been pointed out as playing an important role in mediating inflammation but also as a contributor to its perpetuation, contributing to the pathophysiology of several diseases. In AD, the production of reactive oxygen species (ROS), such as superoxide anion radical (O2•−) and hydroxyl radical (OH) during the oxidative phosphorylation of ATP, exceeds the antioxidant capacity of the biological environment, causing oxidative damage in biomolecules, such as DNA, lipids, and proteins, also leading to neuronal dysfunction and ultimately death [15].
There are currently no commercially available drugs that can stop the progression of AD, only therapies that can ameliorate the symptoms [16]. The available drugs for the treatment of AD act as cholinesterase inhibitors, NMDAR antagonists, and anti-Aβ antibodies. Donepezil, galantamine, and rivastigmine are cholinesterase inhibitors that were found to improve cognition [9,17,18], especially when used as adjunctive with the NMDAR antagonist memantine [19]. Aβ antibodies are gaining more relevance in the treatment of AD, with lecanemab, aducanumab, and donanemab being recently approved by the Food and Drug Administration (FDA). These antibodies, when used in prodromal AD cases, reduce cognitive decline caused by senile plaque formation and improve memory over the span of six months [20,21]. Despite the beneficial effects of the described anti-AD pharmaceuticals, they tend to cause harsh adverse effects, ranging from general gastrointestinal imbalance, abdominal pain, stomach ulcers, and gastritis to confusion, insomnia, hallucinations, arrhythmia, and convulsions [22,23,24,25]. To reduce inflammatory responses by the brain cells, nonsteroid anti-inflammatory drugs, such as ibuprofen, can be used. These can reduce Aβ formation with minor side effects, such as stomach aches and kidney problems [26]. Unfortunately, due to a lack of human testing and general information, there are currently no sufficient data to support nonsteroid anti-inflammatory drugs use in AD patients [27].
Natural products have been used as traditional medicines for centuries. The scientific community is taking advantage of ancient knowledge by investigating the potential protective effect of natural compounds against AD pathogenesis. For instance, the neolignan 4′-O-methylhonokiol, a constituent of the bark of Magnolia officinalis Rehder & E. H. Wilson, was found to inhibit astrocyte activation in preselin-2 mutant (PS2) mice brains and other signaling cascades involved in inflammation and oxidative stress, diminishing memory impairment [28]. Caffeic acid, a phenolic compound naturally present in plants, was shown to have neuromodulatory activity via the regulation of LPS-induced neuroinflammation through tumor necrosis factor-α reduction [29]. The protective effect of caffeic acid against AD was assessed in vivo in a high-fat hyperinsulinemic-induced rat model. The administration of 30 mg/kg of caffeic acid for 30 weeks was shown to ameliorate the memory of the animals, as well as affected the major biochemical targets of AD: decreased the expression of phosphorylated (p)-tau protein, attenuated APP expression, and regulated Aβ deposition [30]. Moreover, the compound 2,2′,4′-trihydroxychalcone isolated from Glycyrrhiza glabra L. was administrated to a double transgenic mouse model used to study early onset AD. The model APP/PS1 mutant mice express a chimeric mouse/human amyloid precursor protein and a mutant variant of human presenilin 1, both connected to the central nervous system. The compound administered at a 9 mg/kg/day (total treatment 106 days) dosage decreased Aβ production and senile plaque formation, which resulted in memory improvement [31]. Natural compounds with antioxidant properties were shown to prevent free radical formation, thus having a potential neuroprotective effect. The most common antioxidants used in AD treatment are resveratrol, apocynin, and other flavonoids which prevent the inactivation of acetylcholine receptors [32,33,34].
A major challenge regarding AD is, in fact, the search for efficient therapies. Presently, most researchers and companies are searching for strategies to discover new targets and new drugs for therapy. The application of natural compounds is among the new features for therapies against AD, as described in recent reviews [25,35,36]. In this field, the potential role of mushrooms for AD treatment or prevention, has been highlighted [37]. Studies using mushroom extracts and compounds have demonstrated neuroprotective, neuromodulatory, and immunomodulatory activities, as well as an anti-inflammatory and antioxidant potential in neurodegenerative diseases [37,38,39]. Well know mushroom species such as Ganoderma lucidum (Curtis) P. Karst., Cordyceps militaris (L.) Fr. and Cyathus africanus H.J. Brodie have been extensively studied with reference to compounds such as ganoderic acid A (3), cordycepin (13), and cyathins (5, 6), respectively (Figure 1). Although there are several studies in the literature focused on the potential of mushroom extracts and compounds in AD, to the best of our knowledge, there is no publication that specifically compiles information on this subject. Considering the pertinence of the research that links mushrooms to AD, we aim with this review to gather knowledge on the application of mushrooms in AD by compiling the existing data to date.

2. Mushrooms in Alzheimer’s Disease

The number of articles linking mushrooms to AD prevention and/or treatment has increased, especially in the last decade. In fact, it was found that mushroom extracts and compounds were efficient against the main anti-AD-related mechanisms, including the inhibition of BACE1 [10,12], the inhibition of the Aβ aggregation, and the inhibition of AChE and butyrylcholinesterase (BChE) (Figure 1). These can be combined with complementary therapeutic approaches, such as anti-inflammatory or antioxidant therapies, which are known to play an important role in neurodegeneration [13,14,26,40]. Mushroom compounds with anti-AD potential are classified into several classes, including lanostanes, terpenoids, indole alkaloids, purine nucleosides, amino acids, lactones, benzaldehyde derivates, dienamides, and ethanolamines as presented in Figure 2 and Figure 3.

2.1. Inhibition of BACE1 and Prevention of Aβ Aggregation and Aβ Cytotoxicity

It has been demonstrated that several mushroom extracts and isolated compounds have anti-BACE1 and anti-Aβ potential (Table 1). Aβ peptide is formed through a cleavage process of APP by the secretases beta (BACE1) and gamma. Aβ peptides tend to interact with each other forming deposits that lead to the formation of amyloid plaques (also referred to as senile plaques), which are considered a hallmark of AD. An excess of Aβ peptide is secreted and accumulates near the synapse. Aβ-dependent neuronal death is thought to be induced due to changes in membrane fluidity and integrity, as these oligomers change the phosphatase and kinase activity, possibly contributing to the hyperphosphorylation of Tau protein. BACE1 has been studied and is considered a promising therapeutic target, as it is involved in Aβ formation [41].
Considering the inhibition of BACE1 and the prevention of Aβ aggregation and Aβ cytotoxicity, species such as Cordyceps militaris, Ganoderma lucidum, Hericium erinaceus (Bull.) Pers. and Pleurotus ostreatus (Jacq.) P. Kumm. are the most representative ones. The mushroom C. militaris has been considered one of the oldest sources of bioactive compounds [42]. The first report on the isolation and characterization of C. militaris metabolites dates to the 1950s and is related to the purine nucleoside cordycepin (13) (Figure 2) [43]. Since then, this molecule has been tested for multiple different purposes, including neurodegenerative diseases, as it exhibits neuroprotective, anti-inflammatory, and immunomodulatory activities [44].
The potential of cordycepin (13) for AD has been evaluated in vivo using an Aβ1–42-induced mouse model where mice were orally administered 100 and 200 mg/kg/day of cordycepin (13) for 15 days [45]. The animals treated with this compound displayed improved cognition, spatial memory, and general behavior (similar results were observed for both concentrations). Rat hippocampal neurons were used as an in vitro model to test cordycepin (13) anti-Aβ potential. The results showed that this compound can inhibit Aβ25–35 production, as well as Aβ-induced ROS and cytotoxicity, p-tau expression, and AChE activity [46].
The medicinal fungus G. lucidum has been used in many countries for health promotion [47]. Wang et al. [48] conducted a study to assess the potential effects of diet supplementation with 0.3, 0.6, and 1.8% of G. lucidum in senescence-accelerated mice prone 8 (SAMP8). Their results showed no significant differences in body weight, eating habits, or locomotion between the treated and control groups (casein diet). However, a significant difference was observed in the behavioral tests, since the supplemented group showed quicker responses in avoidance tests. A significantly higher SOD, GSH-Px, and GSH-Rd activity, as well as the lower accumulation of Aβ was found in the test groups, showing that G. lucidum could be an interesting source of anti-Aβ compounds. Moreover, triterpenoids isolated from G. lucidum were discovered to reduce neuronal apoptosis by inhibiting the ROCK-signaling pathway using APP/PS1 mice [49].
Commonly known as Lion’s Mane Mushroom, H. erinaceus is frequently used in traditional medicine, due to several health benefits such as anti-inflammatory, antibacterial or anticancer properties [50]. H. erinaceus extracts are commercialized as dietary supplements. Several bioactive molecules have been characterized being the most relevant for neurodegeneration erinacine A (7) and S (10), hericenones (1921) (Figure 3), and dilinoleoyl phosphatidylethanolamine (22). The protective effects of erinacines A (7), C (9), and S (10) were tested in in vivo using the double transgenic mouse model APP/PS1. The animals were treated with ethanolic mycelium extracts of H. erinaceus containing these terpenes at 300 mg/kg/day for 30 days (pathological experimental group) and 70–90 days (behavioral experimental group) [51]. The pathological experimental group showed that the treatment schedule led to an inhibition of Aβ deposition, smaller plaque formation, improvement in the Aβ-degrading enzyme neprilysin, enhancement in nerve growth factor (NGF) production, and a promotion in hippocampal neurogenesis, when compared to control. The mice from the behavioral experimental group showed an improvement in daily life tasks, exhibiting higher nesting scores and nestlet shredding. A similar study conducted in SAMP8 fed with erinacine A (7)-enriched mycelium, showed a decrease in Aβ aggregation with an improvement in cognition, memory, and delay in cases of age-related cognitive decline [52]. Tzeng [53] tested the in vivo efficiency of erinacines A (7) and S (10) in APP/PS1 mice. The results showed that both compounds, at concentrations of 10 and 30 mg/kg/day (100 days) reduced Aβ plaque formation, improved plaque degradation, reduced glial activation, and promoted neurogenesis in the highest concentration. Erinacines A (7) displayed the most promising results, with a decrease in Aβ accumulation by inhibiting Aβ production. Behavioral tests also showed an improvement in burrowing and nesting behavior, as well as an increase in spatial learning and memory.
Table 1. Summary of mushroom species, extracts, and/or isolated compounds with anti-BACE1 and anti-Aβ aggregation activity tested in different experimental models.
Table 1. Summary of mushroom species, extracts, and/or isolated compounds with anti-BACE1 and anti-Aβ aggregation activity tested in different experimental models.
SpeciesBody PartExtract/
Compound
In Vitro AssaysIn Vivo AssaysMechanismRefs.
Aleurodiscus vitellinus
(Lév.) Pat.
Cultured myceliumFomannoxinCytotoxicity by MTT;
Microfluorimetry of cytosolic Ca2+; Immunofluorescence for hippocampal neurons; Thioflavin T binding assay for Aβ aggregation
↓ Aβ-induced
cytotoxicity;
preservation of synaptic function in hippocampal neurons;
↓ Aβ binding to neurons
[54]
Amanita caesarea
(Scop.) Pers.
SporocarpAqueous extractHT22 neuron cells:
Apoptosis (annexin V, propidium iodide); MTT; Intracellular ROS; Mitochondrial transmembrane potential (JC-1 staining); Aβ1–42 concentration, SOD, ROS levels by ELISA
AlCl3, D-galactose induced BALB/c mice:
Morris water maze; Rotarod test; Locomotor activity test; Aβ1–42 in serum and brain of BALB/c mice measured by ELISA
↓ Aβ1–42 in brain;
↑ Aβ1–42 in serum
[55]
Antrodia camphorate
(M. Zang & C.H. Su) Sheng H. Wu, Ryvarden & T.T. Chang
Mycelium and
Fruiting body
95% ethanolic extractPC12 cells:
Cytotoxicity by MTT; oxidative stress by DPPH
40 induced Wistar rats:
Morris water maze; Immunoblotting (Aβ and BACE1); IHC—Aβ40 in hippocampus
↓ Aβ-induced cytotoxicity;
↓ Aβ-induced oxidative stress
[56]
Armillaria mellea
(Vahl) P. Kumm.
MyceliumPolysaccharides (precipitated in 70% ethanol) D-galactose- (D-gal-) induced AD mouse model:
Aβ in serum and hippocampus measured by ELISA
↓ Aβ1–42 deposition[57]
Auricularia polytricha
(Mont.) Sacc.
WholeHPLC fractions—aqueous, 100% methanol and 10% ethanol BACE1 inhibition ↓ BACE1;[46,58]
Cordyceps militaris
(L.) Fr.
Cordycepin (13)Hippocampal neurons from rats:
Acridine orange staining (Aβ25–35); Aβ cytotoxicity by MTT; ROS induction
↓ Aβ25–35;
↓ Aβ-induced ROS;
↓ Aβ-induced cytotoxicity
[46]
Cortinarius infractus
(Pers.) Fr.
Fruiting bodyInfractopicrin (11)
10-hydroxy-infractopicrin (12)
HepG2 and SHSY5 cell lines:
Cytotoxicity (CytoTox-One assay); AChE inhibition; Thioflavin T binding assay for Aβ aggregation
↓ Aβ1–40 aggregation in the absence of AChE[59]
Cyathus africanus
H.J. Brodie
Mycelium(12R)-11α,14α-epoxy-13α,14β,15-trihydroxycyathene
Allocyafrin B4
Cyathin I
Cyathin O
BV2 microglial cells:
Cytotoxicity by MTT; iNOS induction by Aβ1–42
↓ Aβ1–42 synthesis;
↓ Aβ-induced iNOS
[60]
Ganoderma lucidum
(Curtis) P. Karst.
Fruiting bodyAqueousNeurons collected from embryonic Sprague Dawley rats: Caspase-like activity; TUNEL staining; Immunofluorescence staining for synaptophysin and Aβ-induced cytotoxicity ↓ Aβ-induced synaptotoxicity by preserving the synaptic density protein, synaptophysin[61]
WholeBiomass SAMP8:
Open field test
Active avoidance test Aging scale
↓ Aβ plaque formation [48]
Triterpenoids APP/PS1 mice: Hematoxylin–eosin staining; IHC; Morris water maze; Exploration test; Burrowing test; TUNEL staining for neuronal apoptosis; Western-blot for Bax, caspase-3, Nrf2, hemeoxidase, NQO1, ROCK-2, GAPDH; SOD; MDA; LDH; Flow cytometry for apoptosis;↓inhibitory effect of Aβ25–35 on neuron proliferation[49]
Hericium erinaceus
(Bull.) Pers.
Cultured mycelia90% Ethanolic extracts containing erinacine A (7), erinacine C (9), and erinacine S (10) APPswe/PS1ΔE9 double transgenic mouse model (APP/PS1) of AD:
Immunoblot; Thioflavin S and AB10 staining for plaque formation and size; ELISA for Aβ levels; Nesting test
↓Aβ deposition;
↓ plaque size;
↑ Aβ-degrading enzymes IDE and neprilysin;
↑NGF
[52]
WholePowdered AlCl3-induced Wistar rats:
Morris water maze; Elevated plus maze; Novel object recognition; IHC; Western blot
↓ Aβ1–42;
↓ APP
[62]
MyceliaErinacine A (7)
Erinacine H
APP/PS1 or SAMP 8 mice:
Thioflavin S and AB10 staining for plaque formation and size; ELISA for Aβ; IHC; Immunostaining for Iba-1, AB10-P, and glial fibrillary acid protein; Nesting test; Burrowing test; Morris water maze
↓Aβ deposition inhibition;
↓ microglial activation;
↑hippocampal neurogenesis
↑dendritic complexity;
↓ Aβ-induced iNOS
[53]
MyceliumErinacine A (7)
enriched mycelium
SAMP8:
Avoidance test; iNOS; TBARS; Aβ-aggregation;
8 OHdG for DNA damage
↓ iNOS;
↓ Aβ aggregation
[63]
Pleurotus ostreatus
(Jacq.) P. Kumm.
WholePolysaccharides Wistar AlCl3/D-galactose induced:
Morris water maze; Step-down memory test; Western blot for APP, BACE1, Aβ, and, PP2A in hippocampus tissue
↓ Aβ formation;
↓ APP;
↓ BACE1;
↑PP2-A
[64]
Ergothioneine
(14)
HeLa cells:
ROS measurement
5XFAD mice
Elevated zero maze; Locomotor activity; Rotarod; Novel object recognition; Fear conditioning; IHC; Dynamic PET imaging
↓ Aβ formation[65]
8-OHdG—8-hydroxy-2′-deoxyguanosine; APP—amyloid precursor protein; BACE1—beta-secretase 1; DPPH—2,2-diphenylpicrylhydrazyl; ELISA—enzyme-linked immunosorbent assay; GAPDH—glyceraldehyde-3-phosphate dehydrogenase; IDE—insulin-degrading enzyme; IHC—immunohistochemistry; iNOS—inducible nitric oxide synthase; LDH—lactate dehydrogenase; MDA—malondialdehyde; MTT-(3-[4,5-dimethylthiazol-2-yl]-2,5 diphenyl tetrazolium bromide; NQO1—NADPH quinone oxiredutase; PP2A—protein phosphatase A; ROCK2—rho-associated protein kinase 2; ROS—reactive oxygen species; SAMP8—senescence-accelerated mouse prone 8; SOD—superoxide dismutase; TBARS—thiobarbituric acid reactive substances; TrkA—tropomyosin receptor kinase A; TUNEL—terminal deoxynucleotidyl transferase dUT; ↑ increase; ↓ decrease.
The species P. ostreatus is the second most popular and most cultivated edible mushroom worldwide. Besides being appreciated for its organoleptic properties, P. ostreatus has been found to induce different bioactivities, such as antimicrobial, antitumor, antioxidant, and immunomodulatory activities [66]. Zhang et al. [64] tested the potential of P. ostreatus polysaccharides to prevent and treat AD. The results of this study showed that administration of 400 mg/kg of polysaccharides extract inhibited BACE1 and Aβ deposition in over 50%, as well as causing a significant inhibition of APP, AChE, and p-tau protein in AlCl3- and D-galactose-induced AD Wistar rats, in which aluminum works as a neurotoxic and D-galactose as a senescence stimulating agent, inducing AD-like symptoms. These improvements were studied through behavior tests and revealed that rats treated with P. ostreatus polysaccharides had a significant improvement in memory and navigation. Another relevant compound for AD produced by P. ostreatus is ergothioneine (14), this sulfur-containing amino acid is known for its antioxidant and cytoprotectant properties. Ergothioneine (14) was tested in the 5XFAD AD transgenic mouse model, characterized for overexpressing five of the known familial AD mutations related to presenilins 1 and 2 that lead to an increased Aβ deposition, higher plasma, and brain levels of Aβ40 and Aβ42, along with elevated pro-inflammatory cytokine levels [67]. Results showed ergothioneine (14) to have metal chelating activity by capturing heavy or potentially toxic metals in the blood stream and removing them from circulation, as well as ROS scavenging activities and the mitigation of Aβ aggregation, which may explain the improvement of the cognitive deficits observed in the experimental group compared to untreated controls [65].

2.2. AChE and BChE Inhibition

One of the most common approaches in AD treatment has been based on repairing cholinergic dysfunction. Cholinergic neurons are involved in several cognitive functions such as memory, attention, learning, and sleep regulation. AD patients have dysfunction in these neurons, mostly in the hippocampus and the neocortex. Damaged cholinergic neurons lead to a decrease in the production of acetylcholine (ACh), a neurotransmitter involved in synapses and general cognition. By lowering the activity of enzymes involved in the degradation of ACh, such as AChE and BChE, the levels of available ACh in the synaptic cleft increase [9,68,69]. When submitted to treatment with ChEIs such as galantamine, a natural compound isolated from Galanthus nivalis L., mice showed an improvement in day-to-day tasks, such as burrowing activities, as well as an enhancement in spatial memory and problem-solving activities [70,71].
Several studies with mushroom extracts and compounds were directed to the anti-AD potential by the evaluation of the AChE and BChE inhibition and the activation of ChAT. In Table 2 the main studies found in the literature are presented.
Ganoderma lucidum aqueous extracts have been shown to be neuroprotective by displaying interesting antioxidant and anti-inflammatory activities in vivo and in vitro, as well as inhibiting AChE activity [72]. Considering compounds, eighteen different lanostane triterpenes were isolated from G. lucidum fruiting bodies. The compounds were tested in vitro and showed AChE inhibition (IC50 ranging between 9.40 and 31.03 µM) [73]. The same study showed that two of the tested compounds, lucidenic acid N (1) and lucidadiol (2) also inhibited BChE, presenting IC50 below 200 µM [73]. This study showed that G. lucidium triterpenes could be potential candidates for anti-AD drugs.
Table 2. Summary of mushroom extracts and/or isolated compounds with AChE and BChE inhibitory activity and/or increased ChAT activity tested in different experimental models.
Table 2. Summary of mushroom extracts and/or isolated compounds with AChE and BChE inhibitory activity and/or increased ChAT activity tested in different experimental models.
SpeciesBody PartExtract/
Compound
In Vitro
Assays
In Vivo
Assays
Refs.
Agaricus
campestris
L.
Fruiting body
Whole
MethanolAChE inhibition; BChE inhibition [74]
Amanita
caesaria
(Scop.) Pers.
SporocarpAqueousEnzyme inhibition of brain and blood ELISA for AChE, ACh, and ChATAlCl3, D-galactose BALB/c mice: Morris water maze; Rotarod test; Locomotor activity test[55]
Amanita
crocea
(Quél.) Singer
Fruiting bodyMethanolic AChE inhibition; BChE inhibition [75,76]
Armillaria
mellea
(Vahl) P. Kumm.
MyceliumPolysaccharides
(precipitated in 70% ethanol)
AlCl3, D-galactose Balb/c mice:
Morris water maze; Fatigue rotarod test; Autonomic activity test; Serum and hypothalamus ACh, AChE, and ChAT; Serum and hypothalamus oxidation status; TUNEL staining; Aβ in serum and hippocampus; IHC
[57]
Cordyceps
militaris
(L.) Fr.
Cordycepin (13)Rat hippocampal neurons:
AChE activity induced by Aβ25–35
[46]
Cortinarius brunneus
(Pers.) Fr.
Fruiting bodyBrunneins A-C
3-(7-hydroxy-9H-β-carbone-1-yl)propanoic acid
AChE inhibition; [77]
Cortinarius
infractus
(Pers.) Fr.
Fruiting bodyInfractopicrin (11)
10-hydroxy-infractopicrin (12)
AChE inhibition; Molecular docking [59]
Cyathus
africanus
H.J. Brodie
Mycelium(12S)-11α,14α-epoxy-13α,14β,15-trihydroxycyath-3-ene (4)
Neocyathin B (5)
Neocyathin J (6)
AChE inhibition; Molecular docking [60,78]
Cyclocybe cylindracea
(DC.) Vizzini & Angelini
Fruiting bodyMethanolicAChE inhibition; BChE inhibition [75]
Ganoderma
lucidum
Fruiting bodyAqueousAChE inhibition [72]
(Curtis) P. Kumm.Fruiting bodyGanoderic acid A (3), B, E, Y
Ganodermadiol
Ganodermanondiol
Ganoderiol F
Lucidadiol (2)
Lucidenic acid A, N (1),
Lucidumol B
Methyl lucidenate E2
Methyl ganoderate A
Methyl ganoderate A acetonide
n-Butyl ganoderate H
n-butyl lucidenate A
n-butyl lucidenate N
AChE inhibition [73]
Hygrocybe acutoconica
(Clem.) Singer
Fruiting bodyAqueousAChE inhibition; BChE inhibition [75]
Inonotus obliquus
(Ach. ex Pers.) Pilát
Lanostante type triterpenoidsAChE inhibition; BchE inhibition;
Molecular docking
[79]
Morchella
esculenta
(L.) Pers.
WholePolysaccharidesAchE inhibition; BchE inhibition [80,81]
Neoboletus erythropus
(Pers.) C. Hahn
Fruiting bodyAqueousAChE inhibition [75]
Phellinus
pini
(Brot.) A. Ames
Fruiting body80% methanolic and hot water extractsAChE inhibition; BChE inhibitionSprague Dawley rat:
Carrageenin-induced hind-paw edema
[82]
Pleurotus
ostreatus
(Jacq.) P. Kumm.
Fruiting bodyAqueousAChE inhibition [75]
Russula
aurea
Pers.
Fruiting bodyAqueousAChE inhibition [75]
Russula
sanguinea
(Bull.) Fr.
Fruiting bodyAqueousAChE inhibition
in solid and liquid
[75]
Trametes
versicolor
(L.) Lloyd
Mycelia and fruiting bodyDMSO, ethanol, or sodium phosphate solution extractsAChE inhibition [83]
Trametes
gibbosa
(Pers.) Fr.
Mycelia and fruiting bodyDMSO, ethanol, or sodium phosphate solution extractsAChE inhibition [83]
Trametes
Hirsute
(Wulfen) Pilát
Mycelia and fruiting bodyDMSO, ethanol, or sodium phosphate solution extractsAChE inhibition [83]
Tremella
fuciformis
Berk.
Fruiting bodyAqueous Scopolamine-treated Sprague Dawley rat:
Morris water maze; IHC for ChAT
[84]
Tricholoma
imbricatum
(Fr.) P. Kumm.
WholeMethanol
Hexane
Ethyl acetate
AChE inhibition; BChE inhibition [85]
ACh—acetylcholine; AChE—acetylcholinesterase; BChE—butyrylcholinesterase; ChAT—choline acetyl transferase; DMSO—dimethyl sulfoxide; IHC—immunohistochemistry; ↑ increase; ↓ decrease.
Amanita caesaria (Scop.) Pers., commonly known as Caesar’s Mushroom, is a well-appreciated edible mushroom commonly consumed in Turkey and is known to have significant antioxidant and antimicrobial activities, as well as a considerable fatty acid composition [86]. A. caesaria aqueous extracts were tested using the immortalized mouse hippocampal cell line HT22 and AlCl3, D-gal induced Balb/c mice [55]. Biochemical examinations found that the extracts promoted cell survival and decreased the expression levels of phosphorylated protein kinase B (p-Akt) and the mammalian target of rapamycin (p-mTOR), known to be associated with cellular death, most efficiently at 50 and 100 µg/mL. The in vivo assays consisted of the intragastrical administration of 250, 500, and 1000 mg/kg/day of extract for 28 days. Vertical movement, locomotor activities, and spatial memory were improved when compared to the AD control group (intragastrically administered saline solution). The inhibition of AChE activity was also observed, as well as an increase in ACh and ChAT. The treatment group was also found to have lower levels of Aβ in the brain and reduced levels of ROS and SOD.
The genera Tricholoma comes from a diverse mushroom family that is composed of edible and non-edible specimens. Several Tricholoma species were found to have antibacterial activity against Gram-negative bacteria [87,88]. In a study developed by Tel et al. [85], the antioxidant and AChE inhibition properties of methanolic, n-hexane, and ethyl-acetate extracts from T. fracticum (Britzelm.) Kreisel, T. terreum (Schaeff.) P. Kumm. and T. imbricatum (Fr.) P. Kumm. were evaluated. Results showed that T. imbricatum hexane extract was the most promising, displaying an inhibitory activity of 71.8% for AChE and 52.6% for BChE at 0.2 mg/mL.
The species Cortinarius infractus (Pers.) Fr. is an inedible mushroom known to have a bitter taste due to its rich composition of infractopicrins (11) [89]. Compounds 11 and 10-hydroxy-infractopicrin (12) showed a significant reduction in AChE activities (IC50 9.72 and 12.7 µM, respectively). The two indole alkaloids also displayed inhibitory activity against Aβ1–40 formation at 25 µM [59].

2.3. Tau Protein Expression and Aggregation

Tau is a protein involved in microtubule formation and function. This protein plays a significant role in cytoplasmatic transport and allows synaptic function as well as the maintenance of their structure, regulating neuronal signaling. The hyperphosphorylation and abnormal cleavage of Tau leads to deficient microtubule activity due to depolymerization consequently causing a loss of neuronal morphology as well as impairments in axon and dendrite formation [6,11]. NFTs are then formed, causing a neuronal malfunction due to impaired synaptic function and neurotoxicity. The level of Tau hyperphosphorylation seems to correlate with the severity of the AD scenario in a direct way [6,11]. In Table 3, the most relevant studies linking mushrooms to protein Tau are compiled followed by a description of the most representative species.
Armillaria mellea (Vahl) P. Kumm. is an edible medicinal mushroom used in traditional medicine in East Asia [90]. Polysaccharides isolated from A. mellea have been reported as having interesting bioactivities, such as antioxidant activity due to radical scavenging potential. A study conducted by An et al. [57] intended to better understand the mechanisms behind A. mellea polysaccharides protection by using an in vitro and an in vivo approach. In vitro, results showed that these molecules improved cell viability in mouse hippocampal neuronal cell line HT22 by inhibiting caspase-3 activity, restoring mitochondrial membrane permeability, and reducing ROS accumulation. As for the in vivo approach, D-galactose-induced Balb/c mice were treated with 25 and 100 mg/kg/day for four weeks. The supplementation with A. mellea polysaccharides regulated AChE, ACh, and ChAT and Aβ concentration in mice serum and hypothalamus (Table 1 and Table 2) as well as a significant reduction in p-tau aggregations in the hippocampus of treated mice at 25 and 100 mg/kg/day.
Table 3. Summary of mushroom extracts and/or compounds with anti-Tau activity tested in different experimental models.
Table 3. Summary of mushroom extracts and/or compounds with anti-Tau activity tested in different experimental models.
SpeciesBody PartExtract/
Compound
In Vitro AssaysIn Vivo AssaysMechanismRefs.
Antrodia camphorata
(M. Zang & C.H. Su) Sheng H. Wu, Ryvarden & T.T. Chang
Mycelium
Fruiting body
Aqueous
Ethanolic
Aβ-induced Wistar rats:
Immunoblotting;
Morris water maze;
↓ p-tau expression[56]
Armillaria
mellea
(Vahl) P. Kumm.
Polysaccharides AlCl3, D-galactose-induced Balb/c mice: Morris water maze; Fatigue rotarod test; Autonomic activity test; TUNEL staining↓ p-tau aggregation;
Antioxidant activity
[57]
Cordyceps
millitaris
(L.) Fr.
Cordycepin (13)Rat hippocampal neurons:
Western blotting for p-tau
↓ Aβ25–35-induced p-tau expression[46]
Hericium
erinaceus
(Bull.) Pers.
WholePowdered AlCl3-induced Wistar rats:
Morris water maze; Elevated plus maze; Novel object recognition; IHC; Western blot
↓ p-tau[62]
Pleurotus
Ostreatus
(Jacq.) P. Kumm.
Whole Polysaccharides Wistar AlCl3/D-galactose-induced rats:
Morris water maze; Step-down memory test; Serum, hippocampus, and liver levels of SOD, GSH-Px, CAT, and MDA; Western blot for p-tau in hippocampus tissue
↓ p-tau[64]
CAT—catalase; GSH-Px—glutathione peroxidase; IHC—immunohistochemistry; iNOS—induced nitric oxide synthase; MDA—malondialdehyde; p-tau—phosphorylated tau; SOD—superoxide dismutase; TUNEL—terminal deoxynucleotidyl transferase dUTP; ↑ increase; ↓ decrease.
As already described, cordycepin (13) isolated from C. militaris was found to inhibit Aβ-induced apoptosis in hippocampal cultivated neurons and to inhibit AChE activity. At 10 µM, the compound also decreased p-tau expression through an adenosine A1 receptor-dependent mechanism.

2.4. Other Activities for General Neuronal Protection

Neuroprotection is generally defined as a physiological or induced measure to preserve neurons. This can be achieved through a therapeutic approach resulting in the salvage, recovery, or even regeneration of the central nervous system cells, structure, and function. Classical examples of neuroprotectors are molecules that intervene in neurotransmitter receptors pathways via agonist/antagonist interaction. These have been explored for their neuroprotective activity as well as their disease-modifying potential. Table 4 summarizes the most relevant mushroom extracts and isolated compounds with neuroprotective potential other than the effects described in the classical AD hypothesis and antioxidative and anti-inflammatory potential.
Grifola frondosa (Dicks.) Gray is an edible mushroom found in America, Europe, and Asia with nutritional and medicinal properties due to its rich composition in bioactive molecules, namely polysaccharides, β-glucans, and heteroglycans [91]. Similarly to members of the Ganoderma genus, G. frondosa is known for its medicinal properties and is used to treat diverse ailments by indigenous Malaysian tribes [92]. An in vivo study by Ling-Sing Seow et al. [93] explored the neuroprotective activities of G. frondosa, G. lucidum, and G. neo-japonicum Imazeki aqueous extracts by testing different concentrations until 2500 ganoderma neo µg/mL in the rat pheochromocytoma cell line PC12, known to have similar behaviors as mature neuronal cells [94]. The results revealed that all extracts stimulated NGF production. While G. frondosa and G. lucidum extracts stimulated NGF production at 12.60 and 12.07% at 75 µg/mL, G. neo-japonicum stimulated NGF production at 14.0% at 50 µg/mL. Lysophosphatidylethanolamine (17), a phospholipid isolated from G. frondosa, was shown to not only stimulate NGF but also enhance neuronal outgrowth and neurofilament M expression while stimulating cell proliferation and survival through the activation of the Ras/MAPK signaling pathway [95].
The lanostane triterpenoid ganodenic acid A (3) isolated from G. lucidum has been shown to inhibit cell apoptosis in brain and liver tissues through the mTOR pathway, while also promoting autophagy, when administered in concentrations of 10 mg/kg/day, when compared to control (treated with normal commercial feed) [96].
Other biologically interesting metabolites produced by H. erinaceus besides erinacines A-C (79) include hericenones and dioleoylphosphatidylethanoleamine (22). Hericenones (1921) are a vast family of benzaldehyde derivates while compound 22 belongs to the phosphatidylethanolamine family. Both are known to have neuroprotective properties [97,98,99]. Hericenones C-E (1921) have been tested in PC12 cells showing the ability to stimulate NGF production through TrkA (tropomyosin kinase A) and MEK (mitogen-activated kinase), as well as neuronal outgrowth mediated by Pi3K/Akt signaling pathway [98,99]. Dilinoleoyl phosphatidylethanolamine (22) extracted from H. erinaceus was tested in Neuro2a cells and showed neuroprotective potential by diminishing endoplasmic reticulum stress-induced death [97]. Termitomyces titanicus Pegler & Piarce-derived compounds, such as 3-(hydroxymethyl)-4-methylfuran-2(5H)-one (15), (3R,4S,1R)-3-(1′-hydroxy-ethyl)-4-methyldihydrofuran-2(3H)-one (16), and 1-hydroxy-3-pentanone (18), are found in Mycoleptonoides aitichisonii Karasaki [100] and the fatty acid amides termitomycamides B (23) and E (24) also displayed neuroprotective activities [101].
Table 4. Summary of mushroom extracts and isolated compounds with neuroprotective activity tested in different experimental models.
Table 4. Summary of mushroom extracts and isolated compounds with neuroprotective activity tested in different experimental models.
SpeciesBody PartExtract/
Compound
In Vitro
Assays
In Vivo
Assays
MechanismRefs.
Armillaria
mellea
(Vahl) P. Kumm.
Mycelium Polysaccharides
(precipitated in 70% ethanol)
HT22 apoptotic cells:
ROS inhibition; MMP depolarization
AlCl3, D-galactose Balb/c mice:
Morris water maze; Fatigue rotarod test; Autonomic activity test; IHC
↑ Cell viability;
↑Behavior
[57]
Cordyceps
millitaris
(L.) Fr.
Cordycepin (13) 1–42 induced ICR mice:
T maze test; Novel object recognition test; Morris water maze; NO scavenging activity
↑Spatial memory
↑Behavior;
↑Memory
[45]
Coriolus
versicolor
(L.) Quél.
WholeBiomass C57BL/6 WT mice: Immunofluorescence; Microscopy for dendritic morphology assessment and β-catenin quantification; Mouse body conditions↑ Arborization of newly generated neurons;
↑ β-catenin
[102]
Dictyophora indusitata
(Vent.) Desv.
Dictyophorine A
Dictyophorine B
Quiescent rat astroglial cells:
NGF production
↑NGF[103]
Ganoderma
lucidum
(Curtis) P. Kumm.
Fruiting body Aqueous Neurons collected from embryonic Sprague Dawley rats:
Aβ-induced cytotoxicity; Caspase-like activity
↓ JNK phosphorylation;
↓ Caspase-3-like activity
[61]
BasidiocarpAqueousPC-12 cells:
Cell viability by MTT; Neurite outgrowth stimulation; Immunofluorescence
↑NGF[93]
Ganodenic acid A (3) C57 BL/6
APP/PS1 mice:
IHC; Western blotting; Transcriptome sequencing; Metabolic analysis
↑p-mTOR;
↓ apoptosis;
↑sphingolipid metabolism;
↑ iron function;
↑autophagy;
Regulation of lipid metabolism;
[96]
Ganoderma neo-japonicum
Imazeki
BasidiocarpAqueousPC-12 cells:
Cell viability by MTT
Neurite outgrowth stimulation
Immunofluorescence
↑NGF[93]
Grifola
frondosa
(Dicks.) Gray
Lysophosphatidylethanolamine (17)PC12 cells:
MAPK activation; Protein phosphorylation; Signal transduction inhibition; Apoptotic DNA fragmentation
↑ neurite outgrowth- GLPE;
↑ neurofilament M expression;
↑Ras/MAPK;
↑NGF
[95]
BasidiocarpAqueousPC-12 cells:
Cell viability by MTT; Neurite outgrowth stimulation; Immunofluorescence
↑NGF[93]
Hericium
erinaceus
(Bull.) Pers.
Basidiocarp Hot water and 80% ethanolic extractNG108-15 (neuroblastoma glial cell hybrid) and MRC-5 (human lung fibroblast) cell lines:
Cell viability by MTT, trypan blue, tunnel assay; Neurite outgrowth stimulation assay; ELISA for NGF
↑NGF[104]
Hericenone C (19),
hericenone D (20),
hericenone E (21)
PC12 cells:
ELISA for NGF; Enzymatic inhibition for MAPK, PI3K, TrkA; Phospho-ERK levels;
Phospho-Akt levels
↑NGF (mediated by TrkA, MEK);
↑ Neurite outgrowth (mediated by Pi3K/Akt)
[99]
Mycelia Erinacine A–C (79)Mouse astroglial cells:
Medium NGF
↑NGF[105]
Erinacine A (7) Wistar rats:
Catecholamine levels (tissue); Indoleamine levels (tissue); ELISA for NGF;
↑NGF;
↑Catecholamine
[106]
dilinoleoyl phosphatidylethanolamine (22)Neuro2a:
Cell viability by MTT; PI staining; Caspase-12 activation; Protein kinase C activation
↓ Endoplasmic reticulum stress-induced death;
↑PKC
[97]
Mycoleptodonoides aitchisonii
Karasaki
Whole Aqueous;
Powder
Rat astrocytes:
NGF
Newborn Wistar rats:
Morris water maze; Monoamine concentration; Amino acid concentration; ELISA for NGF
↑NGF;
↑L-serine;
↑spatial memory
[107]
Terpenoids:
3-(hydroxymethyl)-4-methylfuran-2(5H)-one (15)
(3R,4S,10R)-3-(1′-hydroxy-ethyl)-4-methyldihydrofuran-2(3H)-one (16)
1-hydroxy-3-pentanone (18)
Neuro2a cells:
Cell viability by MTT
↓ Endoplasmic reticulum stress-induced death;
↑NGF
[100]
Paxillus
panuidodes
(Fr.) Fr.
p-terphenyl leucomentins 2–6Mouse cortical cells:
Cell viability by MTT; NMDA cytotoxicity; Glutamate cytotoxicity; Iron chelation—DNA damage
↓ NMDA cytotoxicity;
↓ Glutamate cytotoxicity;
↓DNA damage;
↓ Fenton reactions
[108]
Sarcodon
cyrneus
Maast Geest.
Cyrneine A
Cyrneine B
PC12 and 1321N astrocytoma cells
Medium NGF levels
↑ Neurite growth[109,110]
Sarcodon scabrosus
(Fr.) P. Karst.
Scabronine A1321N astrocytoma cells:
Medium NGF levels; RT-PCR
↑ NFG synthesis[111]
Scabronine B–FRat astroglial cells:
Medium NGF levels
↑ NFG synthesis[101]
Termitomyces albuminosus
(Berk.) R. Heim
Termitomycesphins A–DPC-12 cells:
Neurite growth
↑ Neurite growth[112]
Termitomycesphin E–HPC12 cells
Morphology monitoring—phase contrast microscopy
↑ Neurogenesis [113]
Termitomyces titanicus
Pegler & Piearce
Termitomycamides B (23) and E (24)Neuro2a cells:
Tunicamycin-induced stress test
↓ Endoplasmic reticulum stress-induced death[114]
Tremella
fuciformis
Berk.
Fruiting body Aqueous PC12h cells:
Morphology evaluation by imaging
Scopolamine-treated Sprague Dawley rat:
Morris water maze
↑ neuritogenesis;
↑ NGF;
↑ FGF
[84]
ELISA—enzyme-linked immunosorbent assay; ERK—extracellular-signal-regulated kinase; FGF—fibroblast growth factor; GLPE—thiosulfate transferase; IHC—immunohistochemistry; JNK—jun-amino-terminal kinase; MAPK—mitogen-activated protein kinase; MEK—mitogen-activated kinase; MMP—mitochondrial membrane permeability; MTT- (3-[4,5-dimethylthiazol-2-yl]-2,5 diphenyl tetrazolium bromide; NGF—nerve growth factor; NMDAR—N-methyl-D-aspartate receptor; Nrf2—nuclear factor erythroid 2-related factor 2; p-mTOR—phosphorylated mammalian target of rapamycin; pi3K—phosphoinositide 3-kinase; PKC—protein kinase C; RT-PCR—real time polymerase chain reaction; TrkA—tropomyosin receptor kinase A; ↑ increase; ↓ decrease.
As previously described, inflammatory and oxidative processes have been revealed to be highly deleterious in neurodegeneration scenarios. It is known that mushrooms are composed of many compounds with antioxidant and anti-inflammatory molecules, such as polysaccharides, β-glucans, and phenolic/aromatic compounds. A sizable number of mushroom-isolated compounds display anti-inflammatory and antioxidant activity; therefore, most extracts and compounds described in the previous tables display such activities, thus enhancing the potential of mushroom-derived compounds as possible anti-AD treatment or adjuvants [115]. A recent review by Abitbol et al. [37] efficiently covers the relationship between neuro-inflammation and the potential use of mushroom-derived compounds in AD, for this reason, the topic was not discussed in the present review.
A multi-target therapeutic approach is believed to be more effective due to the multifactorial nature of AD and the fact that some of the hallmark mechanisms are closely related to the point of complementing the pathways involved in other mechanisms [116]. As presented in Table 1, Table 2, Table 3 and Table 4, several mushroom extracts and isolated compounds display one or more anti-AD activities. There are a few examples of extracts displaying more than one activity, including the A. camphorata mycelium and fruiting body 95% ethanolic extract, which reduced Aβ-induced cytotoxicity, oxidative stress, and p-tau expression [56]. G. lucidum fruiting body aqueous extract, was shown to inhibit Aβ-induced cytotoxicity, AChE, and caspase 3-like activity, thereby enhancing cell viability as well as NGF expression [61]. Tremella fuciformis Berk. fruiting body aqueous extract inhibited cholinesterase activity while improving neuritogenesis by stimulating NGF and FGF expression [84]. The most notorious anti-AD mushroom-derived compounds are cordycepin (13) isolated from C. militaris, neocyathins B (5) and J (6) from C. africanus, and erinacine A (7) from H. erinaceus. Cordycepin (13) caused an inhibition in Aβ25–35 production, while reducing Aβ-induced ROS, toxicity, AChE, and tau, resulting in an improvement in spatial memory when administered to mice [46]. Compounds 5 and 6 significantly inhibited Aβ1–42 production, as well as Aβ-induced iNOS and AChE activity [60]. Erinacine A (7) inhibited Aβ deposition, Aβ-induced iNOS, and microglial activation while stimulating neurogenesis through enhanced NGF production [53]. Isolated polysaccharides have also shown interesting combined bioactivities, such as P. ostreatus polysaccharides that displayed neuroprotector activity, inhibiting Aβ formation, APP expression, BACE1 activity, and p-tau deposition [64]. The most interesting results have been found in extracts compared to isolated compounds and compound groups, which can possibly be explained by existing synergistic interactions between compounds in the more diverse extract components.
The results mentioned above highlight the known anti-AD potential exhibited by some extracts and isolated compounds from mushrooms. Although the presented results appear to be interesting, there is still a lack of knowledge about the mechanisms underlying anti-AD protection, which enhances the need for additional research and new isolated molecules that can be applied to anti-AD therapeutics.

2.5. Medicinal Mushrooms in AD Clinical Studies

Mushroom extracts and derivates have been widely accepted as supplements and are thought to improve general health in different ways, such as diminishing chemotherapy side effects; helping with insomnia, anemia, or neurasthenia; and even improving memory [117,118].
Despite the evidence of the anti-AD potential of mushroom extracts and isolated compounds obtained in pre-clinical trials, there is still a lack of published clinical trials.
Some exceptions are clinical trials where patients were treated with Hericium erinaceus. In a double-blind, parallel-group, placebo-control trial that included 30 patients suffering from mild cognitive impairment, 1000 mg of 96% pure mushroom powder was administered three times a day for 16 weeks [119]. Follow-up was performed for 4 weeks after the last dosage and cognitive function was evaluated using a cognitive function scale based on the revised Hasegawa dementia scale. Results showed that, compared to the placebo group, the experimental group exhibited an improvement in cognitive function. Although, at the end of the follow-up, there was a significant decline in results, no major adverse effects were reported with only one participant leaving the study with stomach pain complaints. Additionally, in a more recent study conducted by Li et al. [120] aiming to assess H. erinaceus mycelium enriched with erinacine A (7) potential against early AD, a randomized, double-blind, placebo-controlled pilot study was conducted for 49 weeks. In this trial, patients were administered three capsules daily containing 350 mg of H. erinaceus mycelium and 5 mg/g of compound 7. When compared to the placebo control, the results pointed to an improvement in the instrumental activities of the daily living score as well as Mini-Mental State Examination score, and contrast sensitivity. Concerning AD-specific markers, the experimental group was found to have reduced apolipoprotein APOE4 expression as well as lower concentrations of Aβ1–40 and higher concentrations of brain-derived neurotrophic factor. Four patients left the trial after suffering from abdominal discomfort, nausea, and skin rash. Although not directly connected to AD, during a study concerning the effects of compound 7 enriched H. erinaceus on hearing impairment, Chan et al. [121] found that the administration of 2000 mg/day combined with 10 mg of compound 7 caused a significant increase in serum NGF and brain-derived neurotrophic factor. All of these studies indicate that this medicinal mushroom could be useful as a possible treatment or adjuvant for anti-AD treatment.

3. Conclusions

Given the incidence, severity, and complex nature of Alzheimer’s disease, all efforts made towards finding a cure are critical. Mushrooms are a source of pharmacologically diverse and interesting molecules, produced by different species and body parts. By linking these two subjects, the present review provided an overview of the potential anti-AD properties of mushrooms. Throughout the literature, the link between mushrooms and the disease was obvious and the number of compounds and number of extracts with activity falling within the AD main hypotheses were surprising.
AD is a complex disease with multiple factors involved, which makes it challenging to determine the most crucial mechanisms. However, multi-target inhibitors appear to be the most promising option, as observed in some of the mushroom-derived compounds. These compounds have the potential to be used in anti-AD therapy or as adjuvants in existing treatments. However, although mushroom extracts and derived molecules have been extensively investigated both in vivo and in vitro, currently few clinical trials have been conducted.
Global mushroom production has increased in recent decades and is expected to further increase in the future. As mushrooms are widely consumed all over the world, and their consumption is well accepted when introduced into the market, the knowledge that they can play a positive role in the process of neurodegeneration and cognitive health is valuable and reinforces the necessity of research in this area.

Author Contributions

Conceptualization, A.M.S., L.B. and R.M.; resources, L.B. and R.M.; data curation, A.M.S., C.G., M.R., L.B. and R.M.; writing—original draft preparation, A.M.S. and R.M.; writing—review and editing, M.P., C.G., M.V., V.V., C.D.-M., M.R., L.B. and M.R.; supervision, M.R., L.B. and R.M.; funding acquisition L.B. and R.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work is funded by the European Regional Development Fund (ERDF) through the Regional Operational Program North 2020, within the scope of Project GreenHealth—digital strategies in biological assets to improve well-being and promote green health, Norte-01-0145-FEDER-000042.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Ana Margarida Silva acknowledges the GreenHealth project through the PORTIC/HealthTech/BI/2021/01 grant. Clara Grosso acknowledges the Foundation for Science and Technology (FCT, Portugal) for the grant CEECIND/03436/2020. The authors are also grateful to FCT financial support to CIIMAR (UIDB/04423/2020 and UIDP/04423/2020), CIMO (UIDB/00690/2020 and UIDP/00690/2020) and SusTEC (LA/P/0007/2020) and REQUIMTE/LAQV (UIDB/50006/2020, UIDP/50006/2020 and LA/P/0008/2020); national funding by FCT, P.I., through the institutional scientific employment program contracts of L. Barros.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scheltens, P.; De Strooper, B.; Kivipelto, M.; Holstege, H.; Chételat, G.; Teunissen, C.E.; Cummings, J.; van der Flier, W.M. Alzheimer’s Disease. Lancet 2021, 397, 1577–1590. [Google Scholar] [CrossRef]
  2. Alzheimer Europe. Dementia in Europe Yearbook 2019: Estimating the Prevalence of Dementia in Europe; Alzheimer Europe: Helsinki, Finland, 2019; ISBN 978-99959-995-9-9. [Google Scholar]
  3. Alzheimer’s Association. 2023 Alzheimer’s Disease Facts and Figures. Alzheimers Dement. 2023, 19, 1598–1695. [Google Scholar] [CrossRef] [PubMed]
  4. Maresova, P.; Klimova, B.; Novotny, M.; Kuca, K. Alzheimer’s and Parkinson’s Diseases: Expected Economic Impact on Europe-A Call for a Uniform European Strategy. J. Alzheimers Dis. 2016, 54, 1123–1133. [Google Scholar] [CrossRef]
  5. Alzheimer’s Association. 2020 Alzheimer’s Disease Facts and Figures. Alzheimers Dement. 2020, 16, 391–460. [Google Scholar] [CrossRef] [PubMed]
  6. Iqbal, K.; Liu, F.; Gong, C.-X.; Grundke-Iqbal, I. Tau in Alzheimer Disease and Related Tauopathies. Curr. Alzheimer Res. 2010, 7, 656–664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Murphy, M.P.; Levine, H. Alzheimer’s Disease and the Amyloid-β Peptide. J. Alzheimers Dis. 2010, 19, 311–323. [Google Scholar] [CrossRef]
  8. Bauzon, J.; Lee, G.; Cummings, J. Repurposed Agents in the Alzheimer’s Disease Drug Development Pipeline. Alzheimer’s Res. Ther. 2020, 12, 98. [Google Scholar] [CrossRef]
  9. Sharma, K. Cholinesterase Inhibitors as Alzheimer’s Therapeutics (Review). Mol. Med. Rep. 2019, 20, 1479–1487. [Google Scholar] [CrossRef] [Green Version]
  10. Martins, R.N.; Villemagne, V.; Sohrabi, H.R.; Chatterjee, P.; Shah, T.M.; Verdile, G.; Fraser, P.; Taddei, K.; Gupta, V.B.; Rainey-Smith, S.R.; et al. Alzheimer’s Disease: A Journey from Amyloid Peptides and Oxidative Stress, to Biomarker Technologies and Disease Prevention Strategies-Gains from AIBL and DIAN Cohort Studies. J. Alzheimers Dis. 2018, 62, 965–992. [Google Scholar] [CrossRef] [Green Version]
  11. Metcalfe, M.J.; Figueiredo-Pereira, M.E. Relationship between Tau Pathology and Neuroinflammation in Alzheimer’s Disease. Mt. Sinai J. Med. 2010, 77, 50–58. [Google Scholar] [CrossRef]
  12. Wang, R.; Reddy, P.H. Role of Glutamate and NMDA Receptors in Alzheimer’s Disease. J. Alzheimers Dis. 2017, 57, 1041–1048. [Google Scholar] [CrossRef] [Green Version]
  13. Walker, K.A.; Ficek, B.N.; Westbrook, R. Understanding the Role of Systemic Inflammation in Alzheimer’s Disease. ACS Chem. Neurosci. 2019, 10, 3340–3342. [Google Scholar] [CrossRef] [Green Version]
  14. 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. Alzheimer’s Dement. Transl. Res. Clin. Interv. 2018, 4, 575–590. [Google Scholar] [CrossRef]
  15. Cheignon, C.; Tomas, M.; Bonnefont-Rousselot, D.; Faller, P.; Hureau, C.; Collin, F. Oxidative Stress and the Amyloid Beta Peptide in Alzheimer’s Disease. Redox Biol. 2018, 14, 450–464. [Google Scholar] [CrossRef]
  16. Cummings, J.; Lee, G.; Nahed, P.; Kambar, M.E.Z.N.; Zhong, K.; Fonseca, J.; Taghva, K. Alzheimer’s Disease Drug Development Pipeline: 2022. Alzheimer’s Dement. Transl. Res. Clin. Interv. 2022, 8, e12295. [Google Scholar] [CrossRef]
  17. Hampel, H.; Mesulam, M.M.; Cuello, A.C.; Farlow, M.R.; Giacobini, E.; Grossberg, G.T.; Khachaturian, A.S.; Vergallo, A.; Cavedo, E.; Snyder, P.J.; et al. The Cholinergic System in the Pathophysiology and Treatment of Alzheimer’s Disease. Brain 2018, 141, 1917–1933. [Google Scholar] [CrossRef]
  18. Kokras, N.; Stamouli, E.; Sotiropoulos, I.; Katirtzoglou, E.A.; Siarkos, K.T.; Dalagiorgou, G.; Alexandraki, K.I.; Coulocheri, S.; Piperi, C.; Politis, A.M. Acetyl Cholinesterase Inhibitors and Cell-Derived Peripheral Inflammatory Cytokines in Early Stages of Alzheimer’s Disease. J. Clin. Psychopharmacol. 2018, 38, 138–143. [Google Scholar] [CrossRef]
  19. Liu, J.; Chang, L.; Song, Y.; Li, H.; Wu, Y. The Role of NMDA Receptors in Alzheimer’s Disease. Front. Neurosci. 2019, 13, 43. [Google Scholar] [CrossRef] [Green Version]
  20. Söderberg, L.; Johannesson, M.; Nygren, P.; Laudon, H.; Eriksson, F.; Osswald, G.; Möller, C.; Lannfelt, L. Lecanemab, Aducanumab, and Gantenerumab—Binding Profiles to Different Forms of Amyloid-Beta Might Explain Efficacy and Side Effects in Clinical Trials for Alzheimer’s Disease. Neurotherapeutics 2022, 20, 195–206. [Google Scholar] [CrossRef]
  21. McDade, E.; Cummings, J.L.; Dhadda, S.; Swanson, C.J.; Reyderman, L.; Kanekiyo, M.; Koyama, A.; Irizarry, M.; Kramer, L.D.; Bateman, R.J. Lecanemab in Patients with Early Alzheimer’s Disease: Detailed Results on Biomarker, Cognitive, and Clinical Effects from the Randomized and Open-Label Extension of the Phase 2 Proof-of-Concept Study. Alzheimer’s Res. Ther. 2022, 14, 191. [Google Scholar] [CrossRef]
  22. Withington, C.G.; Turner, R.S. Amyloid-Related Imaging Abnormalities with Anti-Amyloid Antibodies for the Treatment of Dementia Due to Alzheimer’s Disease. Front. Neurol. 2022, 13, 862369. [Google Scholar] [CrossRef] [PubMed]
  23. Imai, H.; Hirai, T.; Kumazawa, R.; Nakagawa, S.; Yonezawa, A.; Matsubara, K.; Nakao, H. Prevalence of and Risk Factors for Adverse Events in Alzheimer’s Patients Receiving Anti-Dementia Drugs in at-Home Care. PLoS ONE 2020, 15, e0231226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Miculas, D.C.; Negru, P.A.; Bungau, S.G.; Behl, T.; Hassan, S.; Tit, D.M. Pharmacotherapy Evolution in Alzheimer’s Disease: Current Framework and Relevant Directions. Cells 2023, 12, 131. [Google Scholar] [CrossRef] [PubMed]
  25. Breijyeh, Z.; Karaman, R. Comprehensive Review on Alzheimer’s Disease: Causes and Treatment. Molecules 2020, 25, 5789. [Google Scholar] [CrossRef] [PubMed]
  26. Lim, G.P.; Yang, F.; Chu, T.; Chen, P.; Beech, W.; Teter, B.; Tran, T.; Ubeda, O.; Ashe, K.H.; Frautschy, S.A.; et al. Ibuprofen Suppresses Plaque Pathology and Inflammation in a Mouse Model for Alzheimer’s Disease. J. Neurosci. 2000, 20, 5709–5714. [Google Scholar] [CrossRef] [Green Version]
  27. Imbimbo, B.P.; Solfrizzi, V.; Panza, F. Are NSAIDs Useful to Treat Alzheimer’s Disease or Mild Cognitive Impairment? Front. Aging Neurosci. 2010, 2, 1517. [Google Scholar] [CrossRef] [Green Version]
  28. Lee, Y.J.; Choi, I.S.; Park, M.H.; Lee, Y.M.; Song, J.K.; Kim, Y.H.; Kim, K.H.; Hwang, D.Y.; Jeong, J.H.; Yun, Y.P.; et al. 4-O-Methylhonokiol Attenuates Memory Impairment in Presenilin 2 Mutant Mice through Reduction of Oxidative Damage and Inactivation of Astrocytes and the ERK Pathway. Free Radic. Biol. Med. 2011, 50, 66–77. [Google Scholar] [CrossRef]
  29. Basu Mallik, S.; Mudgal, J.; Nampoothiri, M.; Hall, S.; Dukie, S.A.; Grant, G.; Rao, C.M.; Arora, D. Caffeic Acid Attenuates Lipopolysaccharide-Induced Sickness Behaviour and Neuroinflammation in Mice. Neurosci. Lett. 2016, 632, 218–223. [Google Scholar] [CrossRef]
  30. 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]
  31. Zhu, Z.; Li, C.; Wang, X.; Yang, Z.; Chen, J.; Hu, L.; Jiang, H.; Shen, X. 2,2’,4’-Trihydroxychalcone from Glycyrrhiza Glabra as a New Specific BACE1 Inhibitor Efficiently Ameliorates Memory Impairment in Mice. J. Neurochem. 2010, 114, 374–385. [Google Scholar] [CrossRef]
  32. ‘t Hart, B.A.; Copray, S.; Philippens, I. Apocynin, a Low Molecular Oral Treatment for Neurodegenerative Disease. Biomed. Res. Int. 2014, 2014, 298020. [Google Scholar] [CrossRef]
  33. Sawda, C.; Moussa, C.; Turner, R.S. Resveratrol for Alzheimer’s Disease. Ann. N. Y. Acad. Sci. 2017, 1403, 142–149. [Google Scholar] [CrossRef]
  34. Khan, H.; Marya; Amin, S.; Kamal, M.A.; Patel, S. Flavonoids as Acetylcholinesterase Inhibitors: Current Therapeutic Standing and Future Prospects. Biomed. Pharmacother. 2018, 101, 860–870. [Google Scholar] [CrossRef]
  35. Lakshmi, S.; Prakash, P.; M Essa, M.; Qoronfleh, W.; Akbar, M.; Song, B.-J.; Kumar, S.; Elumalai, P. Marine Derived Bioactive Compounds for Treatment of Alzheimer’s Disease. Front. Biosci. 2018, 10, 537–548. [Google Scholar]
  36. Wang, S.; Kong, X.; Chen, Z.; Wang, G.; Zhang, J.; Wang, J. Role of Natural Compounds and Target Enzymes in the Treatment of Alzheimer’s Disease. Molecules 2022, 27, 4175. [Google Scholar] [CrossRef]
  37. Abitbol, A.; Mallard, B.; Tiralongo, E.; Tiralongo, J. Mushroom Natural Products in Neurodegenerative Disease Drug Discovery. Cells 2022, 11, 3938. [Google Scholar] [CrossRef]
  38. Lee, W.; Fujihashi, A.; Govindarajulu, M.; Ramesh, S.; Deruiter, J.; Majrashi, M.; Almaghrabi, M.; Nadar, R.M.; Moore, T.; Agrawal, D.C.; et al. Role of Mushrooms in Neurodegenerative Diseases. In Medicinal Mushrooms: Recent Progress in Research and Development; Springer: Singapore, 2019; pp. 223–249. [Google Scholar] [CrossRef]
  39. Dhakal, S.; Kushairi, N.; Phan, C.W.; Adhikari, B.; Sabaratnam, V.; Macreadie, I. Dietary Polyphenols: A Multifactorial Strategy to Target Alzheimer’s Disease. Int. J. Mol. Sci. 2019, 20, 5090. [Google Scholar] [CrossRef] [Green Version]
  40. Khansari, N.; Shakiba, Y.; Mahmoudi, M. Chronic Inflammation and Oxidative Stress as a Major Cause of Age-Related Diseases and Cancer. Recent Pat. Inflamm. Allergy Drug Discov. 2009, 3, 73–80. [Google Scholar] [CrossRef]
  41. Chami, L.; Checler, F. BACE1 Is at the Crossroad of a Toxic Vicious Cycle Involving Cellular Stress and β-Amyloid Production in Alzheimer’s Disease. Mol. Neurodegener. 2012, 7, 52. [Google Scholar] [CrossRef] [Green Version]
  42. Jędrejko, K.J.; Lazur, J.; Muszyńska, B. Cordyceps Militaris: An Overview of Its Chemical Constituents in Relation to Biological Activity. Foods 2021, 10, 2634. [Google Scholar] [CrossRef]
  43. Cunningham, K.G.; Manson, W.; Spring, F.S.; Hutchinson, S.A. Cordycepin, a Metabolic Product Isolated from Cordyceps militaris (Linn.) Link. Nature 1950, 166, 949. [Google Scholar] [CrossRef] [PubMed]
  44. Das, S.K.; Masuda, M.; Sakurai, A.; Sakakibara, M. Medicinal Uses of the Mushroom Cordyceps Militaris: Current State and Prospects. Fitoterapia 2010, 81, 961–968. [Google Scholar] [CrossRef] [PubMed]
  45. He, M.T.; Lee, A.Y.; Kim, J.H.; Park, C.H.; Shin, Y.S.; Cho, E.J. Protective Role of Cordyceps Militaris in Aβ 1–42 -Induced Alzheimer’s Disease in Vivo. Food Sci. Biotechnol. 2019, 28, 865–872. [Google Scholar] [CrossRef] [PubMed]
  46. Song, H.; Huang, L.P.; Li, Y.; Liu, C.; Wang, S.; Meng, W.; Wei, S.; Liu, X.P.; Gong, Y.; Yao, L.H. Neuroprotective Effects of Cordycepin Inhibit Aβ-Induced Apoptosis in Hippocampal Neurons. Neurotoxicology 2018, 68, 73–80. [Google Scholar] [CrossRef] [PubMed]
  47. Cör Andrejč, D.; Knez, Ž.; Knez Marevci, M. Antioxidant, Antibacterial, Antitumor, Antifungal, Antiviral, Anti-Inflammatory, and Nevro-Protective Activity of Ganoderma Lucidum: An Overview. Front. Pharmacol. 2022, 13, 934982. [Google Scholar] [CrossRef]
  48. Wang, M.-F.; Chan, Y.-C.; Wu, C.-L.; Wong, Y.-C.; Hosoda, K.; Yamamoto, S. Effects of Ganoderma on Aging and Learning and Memory Ability in Senescence Accelerated Mice. Int. Congr. Ser. 2004, 1260, 399–404. [Google Scholar] [CrossRef]
  49. Yu, N.; Huang, Y.; Jiang, Y.; Zou, L.; Liu, X.; Liu, S.; Chen, F.; Luo, J.; Zhu, Y. Ganoderma Lucidum Triterpenoids (GLTs) Reduce Neuronal Apoptosis via Inhibition of ROCK Signal Pathway in APP/PS1 Transgenic Alzheimer’s Disease Mice. Oxid. Med. Cell. Longev. 2020, 2020, 9894037. [Google Scholar] [CrossRef]
  50. Ghosh, S.; Nandi, S.; Banerjee, A.; Sarkar, S.; Chakraborty, N.; Acharya, K. Prospecting Medicinal Properties of Lion’s Mane Mushroom. J. Food Biochem. 2021, 45, e13833. [Google Scholar] [CrossRef]
  51. Chen, C.C.; Tzeng, T.T.; Chen, C.C.; Ni, C.L.; Lee, L.Y.; Chen, W.P.; Shiao, Y.J.; Shen, C.C. Erinacine S, a Rare Sesterterpene from the Mycelia of Hericium erinaceus. J. Nat. Prod. 2016, 79, 438–441. [Google Scholar] [CrossRef]
  52. Tsai-Teng, T.; Chin-Chu, C.; Li-Ya, L.; Wan-Ping, C.; Chung-Kuang, L.; Chien-Chang, S.; Chi-Ying, H.F.; Chien-Chih, C.; Shiao, Y.J. Erinacine A-Enriched Hericium erinaceus Mycelium Ameliorates Alzheimer’s Disease-Related Pathologies in APPswe/PS1dE9 Transgenic Mice. J. Biomed. Sci. 2016, 23, 49. [Google Scholar] [CrossRef] [Green Version]
  53. Tzeng, T.T.; Chen, C.C.; Chen, C.C.; Tsay, H.J.; Lee, L.Y.; Chen, W.P.; Shen, C.C.; Shiao, Y.J. The Cyanthin Diterpenoid and Sesterterpene Constituents of Hericium erinaceus Mycelium Ameliorate Alzheimer’s Disease-Related Pathologies in APP/PS1 Transgenic Mice. Int. J. Mol. Sci. 2018, 19, 598. [Google Scholar] [CrossRef] [Green Version]
  54. González-Ramírez, M.; Gavilán, J.; Silva-Grecchi, T.; Cajas-Madriaga, D.; Triviño, S.; Becerra, J.; Saez-Orellana, F.; Pérez, C.; Fuentealba, J. A Natural Benzofuran from the Patagonic Aleurodiscus Vitellinus Fungus Has Potent Neuroprotective Properties on a Cellular Model of Amyloid-β Peptide Toxicity. J. Alzheimers Dis. 2018, 61, 1463–1475. [Google Scholar] [CrossRef]
  55. Li, Z.; Chen, X.; Lu, W.; Zhang, S.; Guan, X.; Li, Z.; Wang, D. Anti-Oxidative Stress Activity Is Essential for Amanita Caesarea Mediated Neuroprotection on Glutamate-Induced Apoptotic HT22 Cells and an Alzheimer’s Disease Mouse Model. Int. J. Mol. Sci. 2017, 18, 1623. [Google Scholar] [CrossRef] [Green Version]
  56. Wang, L.C.; Wang, S.E.; Wang, J.J.; Tsai, T.Y.; Lin, C.H.; Pan, T.M.; Lee, C.L. In Vitro and in Vivo Comparisons of the Effects of the Fruiting Body and Mycelium of Antrodia Camphorata against Amyloid β-Protein-Induced Neurotoxicity and Memory Impairment. Appl. Microbiol. Biotechnol. 2012, 94, 1505–1519. [Google Scholar] [CrossRef]
  57. An, S.; Lu, W.; Zhang, Y.; Yuan, Q.; Wang, D. Pharmacological Basis for Use of Armillaria Mellea Polysaccharides in Alzheimer’s Disease: Antiapoptosis and Antioxidation. Oxid. Med. Cell Longev. 2017, 2017, 4184562. [Google Scholar] [CrossRef] [Green Version]
  58. Bennett, L.; Sheean, P.; Zabaras, D.; Head, R. Heat-Stable Components of Wood Ear Mushroom, Auricularia Polytricha (Higher Basidiomycetes), Inhibit in Vitro Activity of Beta Secretase (BACE1). Int. J. Med. Mushrooms 2013, 15, 233–249. [Google Scholar] [CrossRef]
  59. Geissler, T.; Brandt, W.; Porzel, A.; Schlenzig, D.; Kehlen, A.; Wessjohann, L.; Arnold, N. Acetylcholinesterase Inhibitors from the Toadstool Cortinarius Infractus. Bioorg. Med. Chem. 2010, 18, 2173–2177. [Google Scholar] [CrossRef]
  60. Wei, J.; Cheng, Y.; Guo, W.H.; Wang, D.C.; Zhang, Q.; Li, D.; Rong, J.; Gao, J.M. Molecular Diversity and Potential Anti-Neuroinflammatory Activities of Cyathane Diterpenoids from the Basidiomycete Cyathus Africanus. Sci. Rep. 2017, 7, 8883. [Google Scholar] [CrossRef] [Green Version]
  61. Lai, C.S.W.; Yu, M.S.; Yuen, W.H.; So, K.F.; Zee, S.Y.; Chang, R.C.C. Antagonizing β-Amyloid Peptide Neurotoxicity of the Anti-Aging Fungus Ganoderma lucidum. Brain Res. 2008, 1190, 215–224. [Google Scholar] [CrossRef]
  62. Cordaro, M.; Salinaro, A.T.; Siracusa, R.; D’amico, R.; Impellizzeri, D.; Scuto, M.; Ontario, M.L.; Cuzzocrea, S.; Di Paola, R.; Fusco, R.; et al. Key Mechanisms and Potential Implications of Hericium Erinaceus in NLRP3 Inflammasome Activation by Reactive Oxygen Species during Alzheimer’s Disease. Antioxidants 2021, 10, 1664. [Google Scholar] [CrossRef]
  63. Lee, L.Y.; Chou, W.; Chen, W.P.; Wang, M.F.; Chen, Y.J.; Chen, C.C.; Tung, K.C. Erinacine A-Enriched Hericium erinaceus Mycelium Delays Progression of Age-Related Cognitive Decline in Senescence Accelerated Mouse Prone 8 (Samp8) Mice. Nutrients 2021, 13, 3659. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, Y.; Yang, X.; Jin, G.; Yang, X.; Zhang, Y. Polysaccharides from Pleurotus Ostreatus Alleviate Cognitive Impairment in a Rat Model of Alzheimer’s Disease. Int. J. Biol. Macromol. 2016, 92, 935–941. [Google Scholar] [CrossRef] [PubMed]
  65. Whitmore, C.A.; Haynes, J.R.; Behof, W.J.; Rosenberg, A.J.; Tantawy, M.N.; Hachey, B.C.; Wadzinski, B.E.; Spiller, B.W.; Peterson, T.E.; Paffenroth, K.C.; et al. Longitudinal Consumption of Ergothioneine Reduces Oxidative Stress and Amyloid Plaques and Restores Glucose Metabolism in the 5XFAD Mouse Model of Alzheimer’s Disease. Pharmaceuticals 2022, 15, 742. [Google Scholar] [CrossRef] [PubMed]
  66. Lesa, K.N.; Khandaker, M.U.; Mohammad Rashed Iqbal, F.; Sharma, R.; Islam, F.; Mitra, S.; Emran, T.B. Nutritional Value, Medicinal Importance, and Health-Promoting Effects of Dietary Mushroom (Pleurotus ostreatus). J. Food Qual. 2022, 2022, 2454180. [Google Scholar] [CrossRef]
  67. Oblak, A.L.; Lin, P.B.; Kotredes, K.P.; Pandey, R.S.; Garceau, D.; Williams, H.M.; Uyar, A.; O’Rourke, R.; O’Rourke, S.; Ingraham, C.; et al. Comprehensive Evaluation of the 5XFAD Mouse Model for Preclinical Testing Applications: A MODEL-AD Study. Front. Aging Neurosci. 2021, 13, 713726. [Google Scholar] [CrossRef]
  68. Jasiecki, J.; Wasąg, B. Butyrylcholinesterase Protein Ends in the Pathogenesis of Alzheimer’s Disease—Could BCHE Genotyping Be Helpful in Alzheimer’s Therapy? Biomolecules 2019, 9, 592. [Google Scholar] [CrossRef] [Green Version]
  69. Ferreira-Vieira, T.H.; Guimaraes, I.M.; Silva, F.R.; Ribeiro, F.M. Alzheimer’s Disease: Targeting the Cholinergic System. Curr. Neuropharmacol. 2016, 14, 101–115. [Google Scholar] [CrossRef] [Green Version]
  70. Liu, Y.; Zhang, Y.; Zheng, X.; Fang, T.; Yang, X.; Luo, X.; Guo, A.; Newell, K.A.; Huang, X.F.; Yu, Y. Galantamine Improves Cognition, Hippocampal Inflammation, and Synaptic Plasticity Impairments Induced by Lipopolysaccharide in Mice. J. Neuroinflamm. 2018, 15, 112. [Google Scholar] [CrossRef] [Green Version]
  71. Bhattacharya, S.; Haertel, C.; Maelicke, A.; Montag, D. Galantamine Slows down Plaque Formation and Behavioral Decline in the 5XFAD Mouse Model of Alzheimer’s Disease. PLoS ONE 2014, 9, e89454. [Google Scholar] [CrossRef]
  72. Hasnat, M.A.; Pervin, M.; Lim, B.O. Acetylcholinesterase Inhibition and in Vitro and in Vivo Antioxidant Activities of Ganoderma Lucidum Grown on Germinated Brown Rice. Molecules 2013, 18, 6663–6678. [Google Scholar] [CrossRef] [Green Version]
  73. Lee, I.S.; Ahn, B.R.; Choi, J.S.; Hattori, M.; Min, B.; Bae, K.H. Selective Cholinesterase Inhibition by Lanostane Triterpenes from Fruiting Bodies of Ganoderma Lucidum. Bioorg. Med. Chem. Lett. 2011, 21, 6603–6607. [Google Scholar] [CrossRef]
  74. Akata, I.; Zengin, G.; Picot, C.M.N.; Mahomoodally, M.F. Enzyme Inhibitory and Antioxidant Properties of Six Mushroom Species from the Agaricaceae Family. S. Afr. J. Bot. 2019, 120, 95–99. [Google Scholar] [CrossRef]
  75. Alkan, S.; Uysal, A.; Kasik, G.; Vlaisavljevic, S.; Berežni, S.; Zengin, G. Chemical Characterization, Antioxidant, Enzyme Inhibition and Antimutagenic Properties of Eight Mushroom Species: A Comparative Study. J. Fungi 2020, 6, 166. [Google Scholar] [CrossRef]
  76. Raquel Leal, A.; Barros, L.; Barreira, J.C.; João Sousa, M.; Martins, A.; Santos-Buelga, C.; Ferreira, I.C.F.R. Portuguese Wild Mushrooms at the “Pharma-Nutrition” Interface: Nutritional Characterization and Antioxidant Properties. Food Res. Int. 2013, 50, 1–9. [Google Scholar] [CrossRef]
  77. Teichert, A.; Schmidt, J.; Porzel, A.; Arnold, N.; Wessjohann, L. Brunneins A-C, β-Carboline Alkaloids from Cortinarius Brunneus. J. Nat. Prod 2007, 70, 1529–1531. [Google Scholar] [CrossRef]
  78. Wei, J.; Guo, W.H.; Cao, C.Y.; Kou, R.W.; Xu, Y.Z.; Górecki, M.; Di Bari, L.; Pescitelli, G.; Gao, J.M. Polyoxygenated Cyathane Diterpenoids from the Mushroom Cyathus Africanus, and Their Neurotrophic and Anti-Neuroinflammatory Activities. Sci. Rep. 2018, 8, 2175. [Google Scholar] [CrossRef] [Green Version]
  79. Wei, Y.M.; Yang, L.; Wang, H.; Cai, C.H.; Chen, Z.B.; Chen, H.Q.; Mei, W.L.; Dai, H.F. Triterpenoids as Bivalent and Dual Inhibitors of Acetylcholinesterase/Butyrylcholinesterase from the Fruiting Bodies of Inonotus obliquus. Phytochemistry 2022, 200, 113182. [Google Scholar] [CrossRef]
  80. Badshah, S.L.; Riaz, A.; Muhammad, A.; Çayan, G.T.; Çayan, F.; Duru, M.E.; Ahmad, N.; Emwas, A.H.; Jaremko, M. Isolation, Characterization, and Medicinal Potential of Polysaccharides of Morchella esculenta. Molecules 2021, 26, 1459. [Google Scholar] [CrossRef]
  81. Heleno, S.A.; Stojković, D.; Barros, L.; Glamočlija, J.; Soković, M.; Martins, A.; Queiroz, M.J.R.P.; Ferreira, I.C.F.R. A Comparative Study of Chemical Composition, Antioxidant and Antimicrobial Properties of Morchella esculenta (L.) Pers. from Portugal and Serbia. Food Res. Int. 2013, 51, 236–243. [Google Scholar] [CrossRef] [Green Version]
  82. Im, K.H.; Nguyen, K.; Kim, J.K.; Choi, J.-H.; Lee, T.S. Evaluation of Anticholinesterase and Inflammation Inhibitory Activity of Medicinal Mushroom Phellinus pini (Basidiomycetes) Fruiting Bodies. Int. J. Med. Mushrooms 2016, 18, 1011–1022. [Google Scholar] [CrossRef]
  83. Knežević, A.; Stajić, M.; Sofrenić, I.; Stanojković, T.; Milovanović, I.; Tešević, V.; Vukojević, J. Antioxidative, Antifungal, Cytotoxic and Antineurodegenerative Activity of Selected Trametes Species from Serbia. PLoS ONE 2018, 13, e0203064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Kim, J.H.; Ha, H.-C.; Lee, M.-S.; Kang, J.-I.; Kim, H.-S.; Lee, S.-Y.; Pyun, K.-H.; Shim, I. Effect of Tremella fuciformis on the Neurite Outgrowth of PC12h Cells and the Improvement of Memory in Rats. Biol. Pharm. Bull. 2007, 30, 708–714. [Google Scholar] [CrossRef] [Green Version]
  85. Tel, G.; Apaydın, M.; Duru, M.E.; Öztürk, M. Antioxidant and Cholinesterase Inhibition Activities of Three Tricholoma Species with Total Phenolic and Flavonoid Contents: The Edible Mushrooms from Anatolia. Food Anal. Methods 2012, 5, 495–504. [Google Scholar] [CrossRef]
  86. Doǧan, H.H.; Akbaş, G. Biological Activity and Fatty Acid Composition of Caesar’s Mushroom. Pharm. Biol. 2013, 51, 863–871. [Google Scholar] [CrossRef] [PubMed]
  87. Ovenden, S.P.B.; Yu, J.; Bernays, J.; Wan, S.S.; Christophidis, L.J.; Sberna, G.; Tait, R.M.; Wildman, H.G.; Lebeller, D.; Platel, D.; et al. Trichomycins A and B: Antibacterial Triterpenes from the New Species Tricholoma sp. AU1. J. Nat. Prod. 2005, 68, 409–412. [Google Scholar] [CrossRef]
  88. Hussain, S.; Ahmed, S.; Sardar, R.; Bukhari, S.; Akrem, A.; Mustafa, S.; Manan, A.; Sajid, M.; Talha Rafiq, M. Edible and Medicinal Mushrooms: That Mitigate the Threatening Diseases. Med. Mycol. 2021, 7, 1–6. [Google Scholar]
  89. Brondz, I.; Høiland, K. Biogenesis of Infractine Alkaloids in Cortinarius Infractus: Importance of 5-Hydroxytriptophane Pathway in Biogenesis of Alkaloids in Mushrooms. In Proceedings of the 5th Conference Nordic Separation Science Society, Tallinn, Estonia, 26–29 January 2010. [Google Scholar]
  90. Muszynska, B.; Sulkowska-Ziaja, K.; Wolkowska, M.; Ekiert, H. Chemical, Pharmacological, and Biological Characterization of the Culinary-Medinical Honey Mushroom, Armillaria mellea (Vahl) P. Kumm. (Agaricomycetideae): A Review. Int. J. Med. Mushrooms 2011, 13, 167–175. [Google Scholar] [CrossRef]
  91. Wu, J.Y.; Siu, K.C.; Geng, P. Bioactive Ingredients and Medicinal Values of Grifola frondosa (Maitake). Foods 2021, 10, 95. [Google Scholar] [CrossRef]
  92. Tan, W.C.; Kuppusamy, U.R.; Phan, C.W.; Tan, Y.S.; Raman, J.; Anuar, A.M.; Sabaratnam, V. Ganoderma Neo-Japonicum Imazeki Revisited: Domestication Study and Antioxidant Properties of Its Basidiocarps and Mycelia. Sci. Rep. 2015, 5, 12515. [Google Scholar] [CrossRef] [Green Version]
  93. Ling-Sing Seow, S.; Naidu, M.; David, P.; Wong, K.H.; Sabaratnam, V. Potentiation of Neuritogenic Activity of Medicinal Mushrooms in Rat Pheochromocytoma Cells. BMC Complement. Altern. Med. 2013, 13, 157. [Google Scholar] [CrossRef] [Green Version]
  94. Wiatrak, B.; Kubis-Kubiak, A.; Piwowar, A.; Barg, E. PC12 Cell Line: Cell Types, Coating of Culture Vessels, Differentiation and Other Culture Conditions. Cells 2020, 9, 958. [Google Scholar] [CrossRef]
  95. Nishina, A.; Kimura, H.; Sekiguchi, A.; Fukumoto, R.H.; Nakajima, S.; Furukawa, S. Lysophosphatidylethanolamine in Grifola Frondosa as a Neurotrophic Activator via Activation of MAPK. J. Lipid. Res. 2006, 47, 1434–1443. [Google Scholar] [CrossRef] [Green Version]
  96. Zeng, M.; Qi, L.; Guo, Y.; Zhu, X.; Tang, X.; Yong, T.; Xie, Y.; Wu, Q.; Zhang, M.; Chen, D. Long-Term Administration of Triterpenoids from Ganoderma lucidum Mitigates Age-Associated Brain Physiological Decline via Regulating Sphingolipid Metabolism and Enhancing Autophagy in Mice. Front. Aging Neurosci. 2021, 13, 628860. [Google Scholar] [CrossRef]
  97. Nagai, K.; Chiba, A.; Nishino, T.; Kubota, T.; Kawagishi, H. Dilinoleoyl-Phosphatidylethanolamine from Hericium Erinaceum Protects against ER Stress-Dependent Neuro2a Cell Death via Protein Kinase C Pathway. J. Nutr. Biochem. 2006, 17, 525–530. [Google Scholar] [CrossRef]
  98. Ma, B.J.; Shen, J.W.; Yu, H.Y.; Ruan, Y.; Wu, T.T.; Zhao, X. Hericenones and Erinacines: Stimulators of Nerve Growth Factor (NGF) Biosynthesis in Hericium erinaceus. Mycology 2010, 1, 92–98. [Google Scholar] [CrossRef] [Green Version]
  99. Kawagishi, H.; Ando, M.; Sakamoto, H.; Yoshida, S.; Ojima, F.; Ishiguro, Y.; Ukai, N.; Furukawa, S. Hericenones C, D and E, Stimulators of Nerve Growth Fact (NGF)-Synthesis, from the Mushroom Hericium erinaceum. Tetrahedron Lett. 1991, 32, 4561–4564. [Google Scholar] [CrossRef]
  100. Choi, J.-H.; Horikawa, M.; Okumura, H.; Kodani, S.; Nagai, K.; Hashizume, D.; Koshino, H.; Kawagishi, H. Endoplasmic Reticulum (ER) Stress Protecting Compounds from the Mushroom Mycoleptodonoides aitchisonii. Tetrahedron 2009, 65, 221–224. [Google Scholar] [CrossRef]
  101. Kita, T.; Takaya, Y.; Oshima, Y.; Ohta, T.; Aizawa, K.; Hirano, T.; Inakuma, T. Scabromines B, C, C, D, E and F, Novel Diterpenoids Showing Stimulating Activity of Nerve Growth Factor-Synthesis, from the Mushroom Sarcodon scabrosus. Pergamon Tetrahedron 1998, 54, 11877–11886. [Google Scholar] [CrossRef]
  102. Ferreiro, E.; Pita, I.R.; Mota, S.I.; Valero, J.; Ferreira, N.R.; Fernandes, T.; Calabrese, V.; Fontes-Ribeiro, C.A.; Pereira, F.C.; Rego, A.C. Coriolus Versicolor Biomass Increases Dendritic Arborization of Newly-Generated Neurons in Mouse Hippocampal Dentate Gyrus. Oncotarget 2018, 9, 32929. [Google Scholar] [CrossRef] [Green Version]
  103. Kawagishi, H.; Ishiyama, D.; Mori, H.; Sakamoto, H.; Ishiguro, Y.; Furukawa, S.; Li, J. Dictyophorines A and B, Two Stimulators of NGF-Synthesis from the Mushroom Dictyophora indusitata. Pergamon 1997, 45, 1203–1205. [Google Scholar]
  104. Lai, P.-L.; Naidu, M.; Sabaratnam, V.; Wong, K.H.; David, R.P.; Kuppusamy, U.R.; Abdullah, N.; Malek, S.N.A. Neurotrophic Properties of the Lion’s Mane Medicinal Mushroom, Hericium erinaceus (Higher Basidiomycetes) from Malaysia. Int. J. Med. Mushrooms 2013, 15, 539–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Kawagishia, H.; Shimadaa, A.; Shiraia, R.; Okamotob, K.; Ojimab, F.; Sakamotob, H.; Ishigurob, Y.; Furukawac, S. Erinacines A, B and C, Strong Stimulators of Nerve Growth Factor (NGF)-Synthesis, from the Mycelia of Hericium erinaceum. Pergamon 1994, 35, 1569–1572. [Google Scholar] [CrossRef]
  106. Shimbo, M.; Kawagishi, H.; Yokogoshi, H. Erinacine A Increases Catecholamine and Nerve Growth Factor Content in the Central Nervous System of Rats. Nutr. Res. 2005, 25, 617–623. [Google Scholar] [CrossRef]
  107. Okuyama, S.; Lam, N.V.; Hatakeyama, T.; Terashima, T.; Yamagata, K.; Yokogoshi, H. Mycoleptodonoides aitchisonii Affects Brain Nerve Growth Factor Concentration in Newborn Rats. Nutr. Neurosci. 2004, 7, 341–349. [Google Scholar] [CrossRef]
  108. Lee, I.-K.; Yun, B.-S.; Kim, J.-P.; Ryoo, I.-J.; Kim, Y.-H.; Yoo, I.-D. Neuroprotective Activity of P-Terphenyl Leucomentins from the Mushroom Paxillus panuoides. Biosci. Biotechnol. Biochem. 2003, 67, 1813–1816. [Google Scholar] [CrossRef]
  109. Obara, Y.; Hoshino, T.; Marcotullio, M.C.; Pagiotti, R.; Nakahata, N. A Novel Cyathane Diterpene, Cyrneine A, Induces Neurite Outgrowth in a Rac1-Dependent Mechanism in PC12 Cells. Life Sci. 2007, 80, 1669–1677. [Google Scholar] [CrossRef]
  110. Marcotullio, M.C.; Pagiotti, R.; Maltese, F.; Oball-Mond Mwankie, G.N.; Hoshino, T.; Obara, Y.; Nakahata, N. Cyathane Diterpenes from Sarcodon cyrneus and Evaluation of Their Activities of Neuritegenesis and Nerve Growth Factor Production. Bioorg. Med. Chem. 2007, 15, 2878–2882. [Google Scholar] [CrossRef]
  111. Ohta, T.; Kita, T.; Kobayashi, N.; Obara, Y.; Nakahata, N.; Ohizumi, Y.; Takaya, Y.; Oshima, Y. Scabronine A, a Novel Diterpenoid Having Potent Inductive Activity of the Nerve Growth Factor Synthesis, Isolated from the Mushroom, Sarcodon scabrosus. Tetrahedron Lett. 1998, 39, 6229–6232. [Google Scholar] [CrossRef]
  112. Qi, J.; Ojika, M.; Sakagami, Y. Termitomycesphins A-D, Novel Neuritogenic Cerebrosides from the Edible Chinese Mushroom Termitomyces albuminosus. Tetrahedron 2000, 56, 5835–5841. [Google Scholar] [CrossRef]
  113. Qu, Y.; Sun, K.; Gao, L.; Sakagami, Y.; Kawagishi, H.; Ojika, M.; Qi, J. Termitomycesphins G and H, Additional Cerebrosides from the Edible Chinese Mushroom Termitomyces albuminosus. Biosci. Biotechnol. Biochem. 2012, 76, 791–793. [Google Scholar] [CrossRef] [Green Version]
  114. Paloi, S.; Kumla, J.; Paloi, B.P.; Srinuanpan, S.; Hoijang, S.; Karunarathna, S.C.; Acharya, K.; Suwannarach, N.; Lumyong, S. Termite Mushrooms (Termitomyces), a Potential Source of Nutrients and Bioactive Compounds Exhibiting Human Health Benefits: A Review. J. Fungi 2023, 90, 112. [Google Scholar] [CrossRef]
  115. Elsayed, E.A.; El Enshasy, H.; Wadaan, M.A.M.; Aziz, R. Mushrooms: A Potential Natural Source of Anti-Inflammatory Compounds for Medical Applications. Mediat. Inflamm. 2014, 2014, 805841. [Google Scholar] [CrossRef] [Green Version]
  116. Ibrahim, M.M.; Gabr, M.T. Multitarget Therapeutic Strategies for Alzheimer’s Disease. Neural Regen. Res. 2019, 14, 437–440. [Google Scholar] [CrossRef]
  117. Yadav, S.K.; Ir, R.; Jeewon, R.; Doble, M.; Hyde, K.D.; Kaliappan, I.; Jeyaraman, R.; Reddi, R.N.; Krishnan, J.; Li, M.; et al. A Mechanistic Review on Medicinal Mushrooms-Derived Bioactive Compounds: Potential Mycotherapy Candidates for Alleviating Neurological Disorders. Planta Medica 2020, 86, 1161–1175. [Google Scholar] [CrossRef]
  118. Phan, C.W.; David, P.; Sabaratnam, V. Edible and Medicinal Mushrooms: Emerging Brain Food for the Mitigation of Neurodegenerative Diseases. J. Med. Food 2017, 20, 1–10. [Google Scholar] [CrossRef]
  119. Mori, K.; Inatomi, S.; Ouchi, K.; Azumi, Y.; Tuchida, T. Improving Effects of the Mushroom Yamabushitake (Hericium erinaceus) on Mild Cognitive Impairment: A Double-Blind Placebo-Controlled Clinical Trial. Phytother. Res. 2009, 23, 367–372. [Google Scholar] [CrossRef]
  120. Li, I.C.; Chang, H.H.; Lin, C.H.; Chen, W.P.; Lu, T.H.; Lee, L.Y.; Chen, Y.W.; Chen, Y.P.; Chen, C.C.; Lin, D.P.C. Prevention of Early Alzheimer’s Disease by Erinacine A-Enriched Hericium erinaceus Mycelia Pilot Double-Blind Placebo-Controlled Study. Front. Aging Neurosci. 2020, 12, 155. [Google Scholar] [CrossRef]
  121. Chan, Y.C.; Lin, T.C.; Chen, C.C.; Lee, L.Y.; Chen, W.P.; Liu, Y.Z.; Hwang, J.H. Effects of Erinacine A-Enriched Hericium Erinaceus on Elderly Hearing-Impaired Patients: A Double-Blind, Randomized, Placebo-Controlled Clinical Trial. J. Funct. Foods 2022, 97, 105220. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the main mushrooms bioactive compounds, producer species and AD disease targets. Created with BioRender.com.
Figure 1. Schematic representation of the main mushrooms bioactive compounds, producer species and AD disease targets. Created with BioRender.com.
Molecules 28 05614 g001
Figure 2. Examples of compounds isolated from mushrooms with potential for AD treatment. These bioactive molecules belong to different classes: lanostane triterpenoids (13), polyoxigenated cyathane diterpenoids (49), sesterterpenoids (10), indole alkaloids (11 and 12), and purine nucleosides (13).
Figure 2. Examples of compounds isolated from mushrooms with potential for AD treatment. These bioactive molecules belong to different classes: lanostane triterpenoids (13), polyoxigenated cyathane diterpenoids (49), sesterterpenoids (10), indole alkaloids (11 and 12), and purine nucleosides (13).
Molecules 28 05614 g002
Figure 3. Examples of compounds isolated from mushrooms with potential for AD treatment. These bioactive molecules belong to different classes: amino acids (14), lactones (15 and 16), benzaldehyde derivates (1921), dienamides (23 and 24), and ethanolamines (17, 18 and 22).
Figure 3. Examples of compounds isolated from mushrooms with potential for AD treatment. These bioactive molecules belong to different classes: amino acids (14), lactones (15 and 16), benzaldehyde derivates (1921), dienamides (23 and 24), and ethanolamines (17, 18 and 22).
Molecules 28 05614 g003
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

Silva, A.M.; Preto, M.; Grosso, C.; Vieira, M.; Delerue-Matos, C.; Vasconcelos, V.; Reis, M.; Barros, L.; Martins, R. Tracing the Path between Mushrooms and Alzheimer’s Disease—A Literature Review. Molecules 2023, 28, 5614. https://doi.org/10.3390/molecules28145614

AMA Style

Silva AM, Preto M, Grosso C, Vieira M, Delerue-Matos C, Vasconcelos V, Reis M, Barros L, Martins R. Tracing the Path between Mushrooms and Alzheimer’s Disease—A Literature Review. Molecules. 2023; 28(14):5614. https://doi.org/10.3390/molecules28145614

Chicago/Turabian Style

Silva, Ana Margarida, Marco Preto, Clara Grosso, Mónica Vieira, Cristina Delerue-Matos, Vitor Vasconcelos, Mariana Reis, Lillian Barros, and Rosário Martins. 2023. "Tracing the Path between Mushrooms and Alzheimer’s Disease—A Literature Review" Molecules 28, no. 14: 5614. https://doi.org/10.3390/molecules28145614

Article Metrics

Back to TopTop