Next Article in Journal
Design, Synthesis, and Antisickling Investigation of a Nitric Oxide-Releasing Prodrug of 5HMF for the Treatment of Sickle Cell Disease
Next Article in Special Issue
Acute Effects of Focused Ultrasound-Induced Blood-Brain Barrier Opening on Anti-Pyroglu3 Abeta Antibody Delivery and Immune Responses
Previous Article in Journal
Clade-Specific Alterations within the HIV-1 Capsid Protein with Implications for Nuclear Translocation
Previous Article in Special Issue
Parkin as a Molecular Bridge Linking Alzheimer’s and Parkinson’s Diseases?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Products from Plants and Algae for Treatment of Alzheimer’s Disease: A Review

1
Faculty of Applied Biosciences and Process Technology, Anhalt University of Applied Sciences, Bernburger Straße 55, 06366 Koethen, Germany
2
Paul Flechsig Institute for Brain Research, University of Leipzig, Liebigstraße 19, 04103 Leipzig, Germany
3
Fraunhofer Institute for Cell Therapy and Immunology, Weinbergweg 22, 06120 Halle, Germany
*
Author to whom correspondence should be addressed.
Biomolecules 2022, 12(5), 694; https://doi.org/10.3390/biom12050694
Submission received: 6 March 2022 / Revised: 5 May 2022 / Accepted: 6 May 2022 / Published: 12 May 2022

Abstract

:
Neurodegenerative disorders including Parkinson’s disease (PD), Huntington’s disease (HD) and the most frequent, Alzheimer’s disease (AD), represent one of the most urgent medical needs worldwide. Despite a significantly developed understanding of disease development and pathology, treatments that stop AD progression are not yet available. The recent approval of sodium oligomannate (GV-971) for AD treatment in China emphasized the potential value of natural products for the treatment of neurodegenerative disorders. Many current clinical studies include the administration of a natural compound as a single and combination treatment. The most prominent mechanisms of action are anti-inflammatory and anti-oxidative activities, thus preserving cellular survival. Here, we review current natural products that are either approved or are in testing for a treatment of neurodegeneration in AD. In addition to the most important compounds of plant origin, we also put special emphasis on compounds from algae, given their neuroprotective activity and their underlying mechanisms of neuroprotection.

1. Introduction

Neurodegenerative diseases are a group of disorders in which neuronal function and survival are seriously affected. Many of these diseases, including Parkinson’s, Huntington’s and Alzheimer’s Disease (AD), are caused by structural changes and the deposition of proteins; therefore, they are also assigned to the group of protein misfolding diseases or amyloidoses [1,2,3]. AD is by far the most common cause of neurodegeneration and dementia. It is estimated that AD currently affects 55 million people worldwide (World-Alzheimer-Report-2021. Available online: https://www.alzint.org/u/World-Alzheimer-Report-2021.pdf, accessed on 4 February 2022). Characteristic symptoms of the disease are progressive memory loss, impaired cognitive function and paranoia. The histopathological hallmarks of AD, extracellular amyloid deposits (“amyloid plaques”), which mainly consist of the peptide Aβ, and intraneuronal neurofibrillary tangles of the hyperphosphorylated protein tau, mainly affect the cerebral cortex and the hippocampus [4,5]. Numerous studies suggest that the disease is initiated by the deposition of Aβ, which starts presumably years or decades before the first symptomatic changes [6]. The slow Aβ deposition triggers a downstream cascade (the amyloid cascade), which involves pathologic tau formation and hyperphosphorylation, widespread neuroinflammation and, finally, neuronal death [7,8]. Although the intense research during the last decades enabled a much better understanding of the crucial events in AD pathogenesis, a curative therapy that halts the progression of the disease is not yet available. Most of the so-called disease-modifying experimental drugs are targeting events of the amyloid cascade such as the generation and aggregation of Aβ and the phosphorylation of tau or the cellular metabolism and energy homeostasis [9]. The drug development in AD is faced with several challenges which has resulted in numerous setbacks in recent years [10]. For instance, the enzymes responsible for Aβ formation also have physiological substrates and functions. This complicates the suppression of amyloid peptide formation without interfering with other proteolytical degradation processes. Prominent examples are the γ-secretase complex and the β-secretase BACE1, which play a role in the formation of Aβ peptides [11,12,13]. Moreover, several reports suggest that Aβ1–40/42 and tau also have physiological functions, which leads one to question whether these represent druggable targets [14,15,16,17]. Also, many of the amyloidogenic proteins are localized in the cell nucleus or cytosol, which makes an effective suppression of the aggregation or the breakdown of the conglomerates, e.g., by antibodies, even more difficult [18]. Third, the efficient passage of the blood-brain barrier is needed and thus the pharmaceuticals are required to meet various physicochemical parameters [19,20]. Hence, methods are currently being examined (e.g., focused ultrasound) to make the blood-brain barrier more permeable [21].
Finally, major factors hampering the development and testing of new drugs are based on the clinical presentation of dementia and the currently available diagnostic biomarkers. AD patients frequently also show the presence of Lewy bodies and thus, significant pathological overlap with patients with dementia with Lewy bodies (DLB). As a result, the clinical testing of new active ingredients does not take place in “pure” Alzheimer’s patient populations. Accordingly, attempts are being made (using imaging methods and genetic analyses, among others) to conduct clinical studies in narrowly defined patient populations at an early stage of the disease [22,23,24]. Previously, numerous approaches were therefore undertaken in patients with a possibly too advanced a disease stage [23,25]. In addition, the available diagnostic biomarkers often do not specifically reflect the neurodegenerative disease or provide enough correlation with the clinical status of the patients. These imponderables could be responsible for the failure of different therapeutic approaches in the clinical phase. As mentioned above, alterations in biomarkers precede the symptoms of the disease [6,26], i.e., the measured value of a biomarker cannot be directly correlated with an effect on cognition. An example of this is the antibody bapineuzumab, which caused a significant change in phospho-tau in CSF in phase 2, but missed clinical endpoints [27].
All of these factors finally led to the numerous failures of disease-modifying drugs in AD clinical trials. The very recent accelerated approval of Aducanumab to treat AD may thus represent a first sign of success. However, the complexity also triggered the intense investigations of other fields, such as drugs from natural sources and nutraceuticals (Table 1). One potential reason is that food supplements may have the status as being generally regarded as safe (GRAS) and thus can be quickly applied in clinical testing, and eventually in combination with experimental drugs. Most of these substances are addressing protective mechanisms to cells by, e.g., anti-oxidative effects. However, there are also compounds in testing which are dedicated to disease-modification by, for example, their influence on immune cells. A prominent example is represented by oligomannate from red algae, which obtained approval for AD therapy in China and is currently being tested in additional clinical trials. Due to the emerging role in clinical testing, this review focuses on the current treatment strategies which are based on natural products. We will review drugs which are currently approved but will put a special emphasis on natural products from algae.
This review is based on the personal databases and knowledge of the authors. The work was completed by a substantial amount of literature search using the databases PubMed, Google scholar and SciFinder. The database search was performed until end of February 2022. Only articles in which an active compound was isolated were considered. The date of publication was not an exclusion criterion.

2. Natural Products from Non-Algal Sources

2.1. Esterase Inhibitors

Galantamine. The advanced stage of AD is characterized by a widespread loss of cholinergic basal forebrain neurons [28]. The inhibition of the cholinesterases acetylcholinesterase (AChE) and butyrylcholinesterase (BChE) leads to an increased acetylcholine level in the brain [29,30].
Galantamine [(4aS,6R,8aS)-5,6,9,10,11,12-Hexahydro-3-methoxy-11-methyl-4aH-[1] benzofuro [3a,3,2-ef] [2]benzazepine-6-ol] (Table 2) was first isolated in 1947 from the common snowdrop Galanthus nivalis [31,32]. Later, it was also isolated from Galanthus woronowii and the red spider lily, Lycoris radiata [32,33,34]. In 1960, it was found that galantamine is an inhibitor of cholinesterase [35]. Due to its activity toward muscle AChE, it was used to treat myopathies, post polio paralytic conditions and neuromuscular blockades after anesthesia [36,37]. In 1977 it was reported that galantamine can reverse the acute anticholinergic syndrome induced by scopolamine [38]. The chemical synthesis of galantamine was upscaled and optimized so that quantities of up to 100 kg could be produced under GMP-conditions in the 1990s [39]. Since 2000, Galantamine has been approved in the USA and Europe for the treatment of the symptoms of AD (for example as Reminyl®). It is a reversible, competitive AChE inhibitor and an allosteric modulator of the nicotinic acetylcholine receptors (nAChRs) [40] modulating the α4β2 and α7 nicotinic receptors [41,42,43]. In Phase III studies, it showed side effects like nausea or vomiting with mild severity, mostly during the dose-escalation phase [44].
Huperzine A. Huperzine A, which is isolated from the Chinese club moss Huperzia serrata, is a specific and reversible AChE inhibitor [45]. It binds more tightly and specifically to AChE compared to other inhibitors such as physostigmine, galantamine, donepezil and tacrine [46,47,48]. The dissociation rate from the enzyme is very low [49,50]. The (+)-huperzine A enantiomer and the (−)-huperzine A enantiomer have similar neuroprotective properties, but the (+)-huperzine A enantiomer is 50-fold less potent in inhibiting AChE in an amyloid-β peptide model of toxicity [51]. In another study, the (+)-huperzine A and (−)-huperzine A showed similar results in protecting cells against Aβ toxicity [52]. The neuroprotective effects of huperzine A are created by its potential to protect cells against hydrogen peroxide, β-amyloid toxicity, glutamate, ischemia and staurosporine-induced cytotoxicity and apoptosis [46,47,48,52]. Toxicological studies in different animal species and clinical trials in China have shown that huperzine A has less cholinergic side effects than other AChE inhibitors [47,53,54,55,56]. The most common side effect of huperzine A is nausea [56]. Also, huperzine A improved the memory of aged subjects and patients with AD [54,56]. It is available as a dietary supplement.
Physostigmine. Physostigmine [(3aR,8aS)-1,3a,8-trimethyl-1H,2H,3H,3aH,8H,8aH-pyrrolo [2,3-b]indol-5-ylN-methylcarbamate] is an alkaloid extracted from Physostigma venenosum or Streptomyces pseudogriseolus [57]. It is the oldest known AChE inhibitor. Physostigmine acts as a pesudosubstrate for BChE and AChE, and the inhibition is the result of a transfer of a carbamate residue onto the active site, which is prone to spontaneous hydrolysis and the recovery of the active enzyme. The inhibition of AChE results in an increased acetylcholine level which leads to stimulation of muscarinic and nicotinic receptors [58,59]. Physostigmine can be used as antidote for the anticholinergic toxicity of antihistamines, atropine, tricyclic antidepressants and phenothiazine [60].
Physostigmine is absorbed in the gastrointestinal tract. The bioavailability ranges between 1–8% [61]. It has a short half-life with a peak plasma concentration after 30 min after oral administration of 2 mg [61,62]. To increase the half-life, the slow release physostigmine salicylate was developed [63,64]. Physostigmine can cause several side effects through indirectly influencing muscarinic receptors which could lead, for example, to nausea, vomiting, diarrhea and abdominal pain and nicotinic receptors which could cause paralysis, muscle twitching and the stimulation of cholinergic receptors in the CNS which could lead to CNS depression [65]. Physostigmine derivatives such as tolserine, eseroline and phenserine were synthesized to improve the short half-life and to prevent side effects. Only phenserine was tested in clinical studies [66].

2.2. Plant Natural Products with Antioxidant and Anti-Inflammatory Efficacy

Ginseng. Extracts of the rhizome of the plant Panax ginseng have been used in Asia for thousands of years to treat different diseases including neurological disorders [67].The extract of the plant has several active compounds, ginsenosides, ginseng polysaccharides, volatile oils, peptides and amino acids [68,69]. There are several ginsenosides identified as useful in the treatment of neurodegenerative disease such as AD, PD and HD. The ginsenoside Rb1, Rg1, Rg2, Rg3, Re and Rh2 and Gintonin showed a beneficial effect on AD symptomatology; Rg1, Re and Rd in PD and Ginseng total saponins and Ginsenosides in HD [70,71,72]. The ginsenosides are classified in two groups: the 20(S)-protopanaxadiol (PPD) group and the 20(S)-protopanaxtriol (PPT) group. Rb1, Rc, Rb2, Rd and Rg3 belong to the 20(S)-protopanaxadiol group, while Rg1, Re, Rg2 and Rh1 belong to the 20(S)-protopanaxtriol group [73]. The chemical structure of the ginsenosides is shown in Table 3. Ginsenosides prevent neuroinflammation and oxidative stress. They also have a positive influence on the brain function by apparently diverse mechanisms [74,75,76,77].
For instance, the ginsenoside Rb1 and Rg1 protects spinal cord neurons from oxidative stress induced by H2O2 and excitotoxicity induced by glutamate and kainic acid with an optimal dose of 20–40 µM [67]. In an AD mouse model, Rg1 showed neuroprotective effects through improved cognition and amyloid pathology, modulation of the amyloid precursor protein process and activation of the hippocampal-dependent protein kinase/hippocampal-respond element-binding protein (PKA/CREB) signalling [78]. The ginsenoside Rb1 has several neuroprotective effects. It promotes neural growth, the expression of growth-promoting kinases and helps prevent their levels from decreasing and has played the role of an antiapoptotic agent after Aβ-induced apoptosis in an AD cell model [79,80]. Furthermore, Rb1 seemed to protect the brain from Aluminium-induced toxicity. It reversed the glycogen synthase kinase 3β and the protein phosphates level and thereby reduced tau phosphorylation [81].
Ginkgo biloba. Ginkgo biloba is the oldest living tree species in the world. The standardized Ginkgo biloba extract (GBE) from the dried leaves has neuroprotective effects and is used for the treatment of memory impairment and dementia [83,84]. GBE contains 6% terpenoids, 24% flavonoid glycosides and 5–10% organic acids [85]. The terpenoids include the ginkgolides A, B, C and J (Table 4). Flavonoids and terpenoids are considered to be the pharmacologically active compounds of GBE [86,87]. GBE was shown to reduce the expression of transgenic human amyloid precursor protein expression in mouse brain [88] and to compensate for changes in brain glucose metabolism induced by streptozotocin treatment in rat brain [89].
There are several studies showing a positive effect of GBE on the cognitive function in elderly and AD patients [90,91,92,93]. However, other studies did not show a significant effect in the prevention or treatment of mild cognitive impairment [94,95]. The contradicting outcomes of the studies may be caused by differing compositions of the GBE. The chemical composition depends on the growth conditions and the preparation of the GBE, which highlights the importance to define the composition of drugs derived from natural sources.

2.3. Others

Curcumin is extracted from the rhizome of the Curcuma species. It is the main compound of the curcuminoids and has shown antioxidant and anti-inflammatory properties [96]. In neurological disorders, curcumin decreased inflammation and ROS. Combined with aerobic yoga, curcumin should improve memory and cognitive function (NCT01811381, Table 1).
The main active compounds in elderberry juice, grape powder and Meganatural-Az grape seed extract are anthocyanins. Anthocyanins have anti-inflammatory and antioxidative properties. In animal models of AD, a neuroprotective activity was observed: anthocyanins extracted from black soybeans reversed d-galactose-, lipopolysaccharide- or Aβ1–42-induced oxidative stress and reduced the ROS level [97,98,99,100]. Other anthocyanins inhibited the Aβ- and oxidative stress-induced GSK-3β hyperactivation and hyperphosphorylation of tau protein [101].
Omega-3 poly unsaturated fatty acids (PUFAs) are known to reduce inflammation and vascular risk factors. They decrease cell adhesion molecules which could be related to cerebral small vessel disease [102]. Cerebral small vessel disease influences the accumulation of white matter hyperintensities that results in cognitive decline [103]. Also, metabolites showed neuroprotective properties. The ethyl ester icosapent ethyl from Eicosapentaenoic acid (EPA), an omega-3 PUFA, improves the synaptic function and reduces inflammation (Table 5).
Rapamycin is a macrolide compound from the bacteria Streptomyces hygroscopicus. It inhibits the T and B cell proliferation and was therefore approved by the US Food and Drug Administration (FDA) to suppress the immune system after organ transplantation [104,105,106]. Rapamycin has been shown to reduce Aβ deposition and pathogenic tau phosphorylation to improve synaptic plasticity and to decrease neuroinflammation in mouse models [107,108,109,110,111,112,113].
Cannabinoids from THC-free cannabidiol (CBD) oil target the behavioural and psychological symptoms of dementia. The cannabinoid CBD may act via different mechanisms (Table 5). Several studies suggest that it may protect against Aβ-induced and microglia-activated neurotoxicity in vitro, prevent hippocampal and cortical neurodegeneration, reduce tau hyperphosphorylation and regulate microglial cell migration [114,115,116,117,118]. Furthermore, CBD showed anti-inflammatory and antioxidant activities [119]. The anti-inflammatory properties may result from the decrease of inducible nitric oxide synthase (iNOS) and interleukin-1β protein expression [120]. The anti-inflammatory and neuroprotective properties were investigated in a rat model [121].
Yangxue qingnao is a traditional Chinese medicine composed of 11 different herbs [122]. It is used to improve the cerebral blood flow and thereby the brain nourishment. In a mouse model of AD, Yangxue qingnao pills improved cognitive deficits and reduced Aβ deposition [122]. They possibly promote the expression of α-secretase and thereby the non-amyloidogenic processing of APP [122].

3. Neuroprotective Algal Metabolites

3.1. Carbohydrates

Sodium oligomannate is a mixture of oligosaccharides obtained by the depolymerization of alginate from marine brown algae, followed by its oxidation to oligosaccharides [123,124] (Table 6). In November 2019, it was conditionally approved for the treatment of mild to moderate AD in China [125]. The patients treated with sodium oligomannate showed significant improvement in ADAS-cog12 score compared to the placebo group in a phase II study, whereby the treated group did not show significantly more adverse reactions than the placebo group [126]. The mechanism of action is not completely understood. Studies in mice suggest that oligomannate might act via decreasing neuroinflammation by remodeling gut microbiota and balancing the amino acid metabolism, especially phenylalanine and isoleucine [124].
For other carbohydrates from algae, little or no data are available from in vivo studies. In general, the available data support the mainly anti-oxidative and anti-inflammatory properties of these compounds. Many of these carbohydrates are sulphated and thus strongly negatively charged compounds. Carbohydrates stabilize the cell structure and are involved in ion exchange mechanisms [127,128]. Sulphated polysaccharides from Porphyra haitanesis exhibited antioxidant activity and inhibited lipid peroxidation in rat liver microsomes [129]. The sulphated carbohydrate porphyran from Porphyra yezoensis showed superoxide anion and hydroxyl radical scavenging activity [130]. Sulphated oligosaccharides from the two green algae Ulva lactuca and Enteromorpha prolifera increased concentrations of glutathione, superoxide dismutase (SOD) and catalase (CAT) [131].
Floridoside (2-O-glycerol-α-d-galactopyranoside) extracted from Laurencia undulata showed anti-inflammatory activity in LPS-stimulated BV-2 microglia cells (Table 6). Floridoside inhibited the production of NO and ROS and downregulated iNOS and COX-2 on the gene and protein level via inhibiting the phosphorylation of p38 and ERK [132]. Alginate-derived oligosaccharides inhibited LPS/Aβ42-induced NO and PGE2 synthesis, the expression of COX-2 and iNOS and cytokine release. They diminished the TLR4 and NF-κB overexpression in microglial BV-2 cells [133]. Fucoidan, a fucose-containing sulphated polysaccharide, inhibited ROS and TNF-α release [134]. It reduces NO, PGE2, COX-2, iNOS, MCP-1, TNF-α and IL-1β in LPS-stimulated murine BV2 microglial cells. Fucoidan also decreased the phosphorylation of Akt, ERK, p38 MAPK and JNK [135].
Seleno-polymannuronate is a seleno-derivate from polymannuronate which was synthesized from polymannuronate and Na2SO3 [136]. Polymannuronate is extracted from edible brown algae. Seleno-polymannuronate decreased the production of NO and PGE2 and the expression of COX-2 and iNOS in LPS-treated primary microglia and astrocytes. Sulphated oligosaccharides from the two green algae Ulva lactuca and Enteromorpha prolifera reduced the levels of IL-6, TNF-α and IFN-γ [131]. κ-Carrageenan oligosaccharides and desulphated derivatives inhibited TNF-α secretion in LPS-activated microglia [137].

3.2. Lipids and Proteins

Besides oligosaccharides, lipids have also been described as potential natural products originating from algae that have neuroprotective properties. Hielscher-Michael at al. showed that sulfolipids, membrane components of the thylakoid membrane of microalgae, inhibit the enzyme glutaminyl cyclase (QC). QCs are involved in the formation of pyroglutamate (pGlu)-modified Aβ peptides, whose formation is related to AD pathology [138,139,140]. QC activity is also related to other disorders such as arthritis [141]. QCs catalyse the intramolecular cyclization of N-terminal L-glutamine and glutamate residues into pyroglutamic acid. The modified Aβ peptides are no longer degradable by aminopeptidase and accumulate in the brain. Hence, the inhibition of QC is a potential strategy for the treatment of AD [142]. Hielscher-Michael et al. discovered that 22 methanolic extracts with a concentration of 0.2 mg/mL from the algae Scenedesmus rubescens, Scenedesmus producto-capitatus, Scenedesmus accuminatus, Scenedesmus pectinatus, Tetradesmus wisconsinensis and Eustigmatos magnus showed QC inhibitory activity between 15% to 72% [143]. The compounds with QC inhibitory activity were identified as the sulfolipids 1,2-di-O-palmitoyl-3-O-(6′-deoxy-6′-sulfo-D-glycopyranosyl)-glycerol, 1-O-palmitoyl-2-O-linolenyl-3-O-(6′-deoxy-6′-sulfo-D-glucopyranosyl)-glycerol and 1-O-linolyl-2-O-palmitoyl-3-O-(6′-deoxy-6′-sulfo-D-glucopyranosyl)-glycerol (Table 7) [143].
The glycoprotein of Undaria pinnatifida (UPGP) has antioxidant properties through the enhancing of superoxide dismutase (SOD) activity and inhibiting xanthine oxidase (Xox) activity at a concentration of 5 mg/mL and 1 mg/mL [146]. UPGP showed anti-inflammatory properties in LPS-stimulated RAW264.7 macrophages via inhibition of COX-1, COX-2 and NO [146]. UPGP has AChE, BChE and BACE1 inhibitory activities [146]. UPGP inhibited the BACE1 activity in in vitro enzymatic assays [146].
The cyanobacterial peptides tasiamide B and its analog tasiamide F, both isolated from the marine cyanobacterium Lyngbya sp., showed BACE-1 (β-site of APP cleaving enzyme) inhibitory activity [144]. Tasiamide B is a more effective inhibitor of BACE-1 [144,145]. It was also extracted from Symploca sp., another marine cyanobacterium [145].

3.3. Phenols

The bioactive and neuroprotective polyphenols have been typically isolated from brown algae. Typically, they interfere with several signal transduction pathways or function as enzyme inhibitors (Table 8). For instance, eckol, dieckol and 8,8′-bieckol from Ecklonia cava showed anti-inflammatory properties in Aβ25–35-stimulated PC12 cells by inhibition of TNF-α, IL-1β and PGE2 synthesis [147]. These phlorotannins further downregulated the proinflammatory enzymes iNOS and COX-2 by interference with the NF-κB pathway [147]. Dieckol suppressed p38, ERK and JNK, while eckol suppressed the activation of p38 and 8,8′-bieckol decreased the phosphorylation of p38 and JNK [147]. In another experiment, dieckol from Ecklonia cava suppressed the production of NO and PGE2 and the expression of iNOS and COX-2 in LPS-stimulated murine BV2 microglia. The reduction of IL-1β, TNF-α, NFκB, p38 and ROS was also shown before by others [148]. Antioxidant properties were also observed with diphlorethohydroxycarmalol and 6,6′-bieckol isolated from Ishige okamurae [149,150].
Phlorofucofuroeckol B isolated from Ecklonia stolonifera lowered the expression of COX-2 and inducible nitric oxide synthase in LPS-stimulated BV-2 cells [151]. It reduced the pro-inflammatory cytokines IL-1β, IL-6 and TNF-α. It prevents the degradation of inhibitor κB-α (IκB-α) and thereby inhibits the activation of NF-κB. Phlorofucofuroeckol B also inhibited the phosphorylation of Akt, ERK and JNK [151]. The phlorotannins phloroglucinol, eckol, dieckol, 7-phloroeckol, phlorofucofuroeckol A and dioxinodehydroeckol from Eisenia bicyclis inhibited NO production [152]. Phlorofucofuroeckol A from Ecklonia stolonifera attenuated NO, PGE2, iNOS and COX-2 expression [153]. It lowers the level of IL-1β, IL-6 and TNF-α. As Phlorofucofuroeckol B, Phlorofucofuroeckol A prevents the degradation of IκB-α and inhibits thereby the activation of NF-κB. Phlorofucofuroeckol A downregulated JNK, p38 and Akt [153]. 8,8′-bieckol reduced ROS, NO, PGE2, IL-6 and iNOS in LPS-stimulated primary macrophages, RAW264.7 macrophages and LPS-induced septic mice. It lowers the transactivation and NF-κB and nuclear translocation of the NF-κB p65 subunit [154]. 6,6′-bieckol from Ecklonia stolonifera attenuated IL-6, NO, PGE2, COX-2 and iNOS in LPS-stimulated BV2 and murine primary microglial cells. It inhibited the transactivation of NF-κB and the nuclear translocation of the NF-κB p65 subunit as well as the phosphorylation of Akt, JNK and p38 MAPK [155]. The phloroglucinol derivatives dibenzo [1,4]dioxine-2,4,7,9-tetratol from Ecklonia maxima inhibited AChE [156], while 6,6′-bieckol extracted from the red algae Grateloupia elliptica inhibited AChE and BChE [157].
Sargachromenol isolated from Sargassum micracanthum decreased NO, PGE2, COX-2 and iNOS and increased IκB-α [158]. Sargaquinoic acid extracted from Sargassum siliquastrum showed anti-inflammatory activity trough reducing NO and iNOS, nuclear translocation of NF-κB and JNK1/2 MAPK. It prevents the degradation of IκB-α [159].
Some polyphenols also showed inhibitory activity on esterases. The phlorotannins phloroglucinol, dibenzo [1,4]dioxine-2,4,7,9-tetraol and eckol showed AChE inhibition in in vitro enzyme assays [156]. Dieckol and phlorofucofuroeckol extracted from Ecklonia cava inhibited AChE and increased the level of acetylcholine in mice [160].
Sargaquinoic acid and sargachromenol isolated from Sargassum sagamianum and Sargassum serratifolium and sargahydroquinic acid extracted from Sargassum serratifolium showed moderate AChE inhibitory properties and BACE-1 inhibitory activity. Sargaquinoic acid is a potent BChE inhibitor [161,162].
The polyphenols eckol, dieckol, phloroglucinol and dioxinodehydroeckol extracted from Ecklonia stolonifera inhibited the self-aggregation of Aβ25–35 in vitro [163].

3.4. Isoprenoids

Similar to polyphenols, the neuroprotective effect of isoprenoids such as sterols and xanthin derivatives is primarily based on their anti-oxidative radical scavenging and anti-inflammatory properties (Table 9). Numerous studies have been published addressing the antioxidative activity in different, mostly cellular model systems. For instance, the steroid fucosterol extracted from Pelvetia siliquosa increased the level of antioxidant enzymes SOD, GPx and CAT and inhibited ROS production [152,168]. It also provided protection from oxidative damage by raising the GSH level and attenuated of the production of iNOS, TNF-α and IL-6, and the phosphorylation of NF-κB, MKK3/6 and MK2 was shown [169,170,171]. Fucosterol from Panida australis and Hizikia fusiformis reduced IL-1β, IL-6, TNF-α, NO and PGE2 in LPS- or Aβ-induced BV2 microglia cells or keratinocytes [172,173]. Fucosterol extracted from the algae Ecklonia stolonifera, Panida australis and Sargassum horridum inhibited AChE and BChE in vitro [172,174,175]. Different types of inhibition were detected depending on the origin. Fucosterol from Ecklonia stolonifera showed a selective inhibition of BChE, a non-selective cholinesterase inhibition of AChE and BChE was observed with fucosterol from Panida australis and a non-competitive inhibition was detected with the compound from Sargassum horridum [172,174,175]. A non-competitive inhibition of the β-secretase BACE1 was observed with fucosterol from Ecklonia stolonifera and Undaria pinnatifida [176].
The carotenoid fucoxanthin extracted from Sargassum siliquastrum prevented H2O2-induced and reduced ROS-induced DNA damage [177,178]. It also decreased the cytokines IL-6, IL-1β, TNF-α, NO and PGE2 and the enzyme activity of COX-2 and iNOS by suppressing the phosphorylation of MAPKs in Aβ42-induced BV-2 microglia cells [177]. In the presence of fucoxanthin, enhanced cell survival was observed with LPS-activated BV-2 microglia by activation of the cAMP-dependent signal cascade pathway resulting in the attenuation of the phosphorylation of Akt, NF-κB, ERK, p38 MAPK and AP-1 and reduced levels of TNF-α, IL-6, PGE2, NO and ROS [179]. Fucoxanthin activated the nuclear factor erythroid 2-related factor 2 (Nrf2)/heme oxygenase-1 (HO-1) pathway and increased the secretion of brain-derived neurotrophic factor [179].
Fucoxanthin isolated from Phaeodactylum tricornutum inhibited BChE activity in vitro [180]. It possibly interacts with a peripheral anionic site of AChE mediating non-competitive inhibition [183]. Similarly, α-Bisabolol isolated from Padina gymnospora inhibited AChE and BChE in vitro [182]. In two other studies, Fucoxanthin suppressed the formation of Aβ1-42 fibrils and oligomers and inhibited Aβ aggregation [184,185]. α-Bisabolol prevents oligomer formation and disaggregates the mature fibrils [186].
Astaxanthin decreased the cytokine levels of IL-6, IL-1β, and TNF-α. It inhibited iNOS, nNOs and COX-2 expression in the hippocampus and prefrontal cortex of male mice [181]. In rats, astaxanthin attenuated NF-κB activity and the expression of IL-1β, TNF-α and the intercellular adhesion molecule 1 [187].

4. Conclusions

The recent conditional approval of the monoclonal antibody aducanumab (aduhelm) by the FDA provides a very stimulating signal for all drug development approaches in AD. However, among others, these antibody approaches are still met with doubts about disease modification and safety, as suggested by the decision of the EMA to not provide approval to Aduhelm (Meeting highlights from the Committee for Medicinal Products for Human Use (CHMP) 13–16 December 2021. Available online: https://www.ema.europa.eu/en/news/meeting-highlights-committee-medicinal-products-human-use-chmp-13-16-december-2021, accessed on 2 February 2022). Hence, nutritional approaches and natural products are vital tools for prevention and amelioration of the progression of neurodegeneration. A considerable strength of the natural products is provided by the multifaceted mechanisms of their activity. Prominent examples for that include, for instance, the ginsenosides or the extracts from Ginkgo biloba (GBE), which are currently the subject of late-stage clinical trials (Table 1). The ingredients exert anti-inflammatory and antioxidative properties and have been described to influence the processing of AD-related proteins, providing a multi-pronged molecular approach of intervention. Also, natural compounds are among the first described to address potential novel pathways in neurodegenerative diseases. The most prominent example for that is GV-971 (sodium oligomannate). The currently available data support an influence on the gut microbiome which leads to the amelioration of AD-related symptomatology. The compound is among the first that addresses the “gut-brain-axis”, which has recently become focus of research in neurodegenerative diseases. Besidessodiumoligomannate, the general role of nutrition and nutrient uptake by the digestive tract is further underscored by the recent reports on the LipiDiDiet multinutrient clinical trial in prodromal Alzheimer’s disease [188]. Collectively, the unique properties of these molecules should further encourage the evaluation of combination therapies of, for example, anti-Aβ immunotherapy and treatment with natural products. Because the compounds reviewed here are mostly available without a prescription, a quick introduction into theclinical routine thus appears straightforward.

Funding

This research received no external funding.

Acknowledgments

We are grateful to Daniel Ramsbeck for critical reading of the manuscript.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Rostagno, A.; Holton, J.L.; Lashley, T.; Revesz, T.; Ghiso, J. Cerebral amyloidosis: Amyloid subunits, mutants and phenotypes. Cell. Mol. Life Sci. 2009, 67, 581–600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sipe, J.D.; Benson, M.D.; Buxbaum, J.N.; Ikeda, S.-I.; Merlini, G.; Saraiva, M.J.; Westermark, P. Amyloid fibril protein nomenclature: 2010 recommendations from the nomenclature committee of the International Society of Amyloidosis. Amyloid 2010, 17, 101–104. [Google Scholar] [CrossRef] [PubMed]
  3. Hazenberg, B.P. Amyloidosis: A Clinical Overview. Rheum. Dis. Clin. N. Am. 2013, 39, 323–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Braak, H.; Braak, E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 1991, 82, 239–259. [Google Scholar] [CrossRef]
  5. Thal, D.R.; Rüb, U.; Orantes, M.; Braak, H. Phases of Aβ-deposition in the human brain and its relevance for the development of AD. Neurology 2002, 58, 1791–1800. [Google Scholar] [CrossRef]
  6. Hadjichrysanthou, C.; Evans, S.; Bajaj, S.; Siakallis, L.C.; McRae-McKee, K.; de Wolf, F.; Anderson, R.M.; Initiative, T.A.D.N. The dynamics of biomarkers across the clinical spectrum of Alzheimer’s disease. Alzheimer’s Res. Ther. 2020, 12, 1102. [Google Scholar] [CrossRef]
  7. Hardy, J.A.; Higgins, G.A. Alzheimer’s disease: The amyloid cascade hypothesis. Science 1992, 256, 184–185. [Google Scholar] [CrossRef]
  8. Uddin, S.; Kabir, T.; Rahman, S.; Behl, T.; Jeandet, P.; Ashraf, G.M.; Najda, A.; Bin-Jumah, M.N.; El-Seedi, H.R.; Abdel-Daim, M.M. Revisiting the Amyloid Cascade Hypothesis: From Anti-Aβ Therapeutics to Auspicious New Ways for Alzheimer’s Disease. Int. J. Mol. Sci. 2020, 21, 5858. [Google Scholar] [CrossRef]
  9. Cummings, J.; Lee, G.; Zhong, K.; Fonseca, J.; Taghva, K. Alzheimer’s disease drug development pipeline. Alzheimer’s Dementia Transl. Res. Clin. Interv. 2021, 7, e12179. [Google Scholar] [CrossRef]
  10. Long, J.M.; Holtzman, D.M. Alzheimer Disease: An Update on Pathobiology and Treatment Strategies. Cell 2019, 179, 312–339. [Google Scholar] [CrossRef]
  11. Lichtenthaler, S.F.; Tschirner, S.K.; Steiner, H. Secretases in Alzheimer’s disease: Novel insights into proteolysis of APP and TREM. Curr. Opin. Neurobiol. 2021, 72, 101–110. [Google Scholar] [CrossRef]
  12. Imbimbo, B.P.; Watling, M. Investigational BACE inhibitors for the treatment of Alzheimer’s disease. Expert Opin. Investig. Drugs 2019, 28, 967–975. [Google Scholar] [CrossRef]
  13. Luo, J.E.; Li, Y.-M. Turning the tide on Alzheimer’s disease: Modulation of γ-secretase. Cell Biosci. 2022, 12, 2. [Google Scholar] [CrossRef]
  14. Puzzo, D.; Privitera, L.; Leznik, E.; Fà, M.; Staniszewski, A.; Palmeri, A.; Arancio, O. Picomolar Amyloid-β Positively Modulates Synaptic Plasticity and Memory in Hippocampus. J. Neurosci. 2008, 28, 14537–14545. [Google Scholar] [CrossRef]
  15. Puzzo, D.; Privitera, L.; Fa’, M.; Staniszewski, A.; Hashimoto, G.; Aziz, F.; Sakurai, M.; Ribe, E.M.; Troy, C.M.; Mercken, M.; et al. Endogenous amyloid-β is necessary for hippocampal synaptic plasticity and memory. Ann. Neurol. 2011, 69, 819–830. [Google Scholar] [CrossRef] [Green Version]
  16. Puzzo, D.; Arancio, O. Amyloid-β Peptide: Dr. Jekyll or Mr. Hyde? J. Alzheimer’s Dis. 2013, 33, S111–S120. [Google Scholar] [CrossRef] [Green Version]
  17. Iqbal, K.; Liu, F.; Gong, C.-X. Tau and neurodegenerative disease: The story so far. Nat. Rev. Neurol. 2016, 12, 15–27. [Google Scholar] [CrossRef]
  18. E Golde, T. Open questions for Alzheimer’s disease immunotherapy. Alzheimer’s Res. Ther. 2014, 6, 3. [Google Scholar] [CrossRef]
  19. Rankovic, Z. CNS Drug Design: Balancing Physicochemical Properties for Optimal Brain Exposure. J. Med. Chem. 2015, 58, 2584–2608. [Google Scholar] [CrossRef]
  20. Hubbard, R.E. Structure-based drug discovery and protein targets in the CNS. Neuropharmacology 2011, 60, 7–23. [Google Scholar] [CrossRef]
  21. Liu, X.; Naomi, S.S.M.; Sharon, W.L.; Russell, E.J. The Applications of Focused Ultrasound (FUS) in Alzheimer’s Disease Treatment: A Systematic Review on Both Animal and Human Studies. Aging Dis. 2021, 12, 1977. [Google Scholar] [CrossRef] [PubMed]
  22. Ringman, J.M.; Network, D.I.A.; Goate, A.; Masters, C.L.; Cairns, N.J.; Danek, A.; Graff-Radford, N.; Ghetti, B.; Morris, J.C. Genetic Heterogeneity in Alzheimer Disease and Implications for Treatment Strategies. Curr. Neurol. Neurosci. Rep. 2014, 14, 499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Schneider, L.S.; Mangialasche, F.; Andreasen, N.; Feldman, H.; Giacobini, E.; Jones, R.; Mantua, V.; Mecocci, P.; Pani, L.; Winblad, B.; et al. Clinical trials and late-stage drug development for Alzheimer’s disease: An appraisal from 1984 to 2014. J. Intern. Med. 2014, 275, 251–283. [Google Scholar] [CrossRef] [PubMed]
  24. Duara, R.; Barker, W. Heterogeneity in Alzheimer’s Disease Diagnosis and Progression Rates: Implications for Therapeutic Trials. Neurotherapeutics 2022, 1–18. [Google Scholar] [CrossRef]
  25. Giacobini, E.; Gold, G. Alzheimer disease therapy—Moving from amyloid-β to tau. Nat. Rev. Neurol. 2013, 9, 677–686. [Google Scholar] [CrossRef]
  26. Bateman, R.J.; Xiong, C.; Benzinger, T.L.S.; Fagan, A.M.; Goate, A.; Fox, N.C.; Marcus, D.S.; Cairns, N.J.; Xie, X.; Blazey, T.M.; et al. Clinical and Biomarker Changes in Dominantly Inherited Alzheimer’s Disease. N. Engl. J. Med. 2012, 367, 795–804. [Google Scholar] [CrossRef] [Green Version]
  27. Zetterberg, H.; Blennow, K.; Rinne, J.O.; Salloway, S.; Wei, J.; Black, R.; Grundman, M.; Liu, E.; for the AAB-001 201/202 Investigators. Effect of Immunotherapy with Bapineuzumab on Cerebrospinal Fluid Biomarker Levels in Patients With Mild to Moderate Alzheimer Disease. Arch. Neurol. 2012, 69, 1002–1010. [Google Scholar] [CrossRef]
  28. Maurer, S.V.; Williams, C.L. The Cholinergic System Modulates Memory and Hippocampal Plasticity via Its Interactions with Non-Neuronal Cells. Front. Immunol. 2017, 8, 1489. [Google Scholar] [CrossRef] [Green Version]
  29. Giacobini, E. Cholinesterase inhibitors: New roles and therapeutic alternatives. Pharmacol. Res. 2004, 50, 433–440. [Google Scholar] [CrossRef]
  30. Folch, J.; Petrov, D.; Ettcheto, M.; Abad, S.; López, E.S.; García, M.L.; Olloquequi, J.; Beas-Zarate, C.; Auladell, C.; Camins, A. Current Research Therapeutic Strategies for Alzheimer’s Disease Treatment. Neural Plast. 2016, 2016, 8501693. [Google Scholar] [CrossRef] [Green Version]
  31. Proskurnina, N.; Areshknina, L. Alkaloids of Galanthus woronovi. Allg. Chem. 1947, 17, 1216–1219. [Google Scholar]
  32. Mucke, H.A. The case of galantamine: Repurposing and late blooming of a cholinergic drug. Future Sci. OA 2015, 1, FSO73. [Google Scholar] [CrossRef]
  33. Proskurnina, N.; Yakovieva, A. Alkaloids of Galanthus woronovi. Structure of galanthine. Ber. Akad. Wiss. UdSSR 1953, 90, 565–567. [Google Scholar]
  34. Uyeo, S.; Kobayashi, S. Lycoris Alkaloids. XXIV.: Isolation and Characterization of Lycoremine. Pharm. Bull. 1953, 1, 139–142. [Google Scholar] [CrossRef] [Green Version]
  35. Irwin, R.; Smith, H. Cholinesterase inhibition by galanthamine and lycoramine. Biochem. Pharmacol. 1960, 3, 147–148. [Google Scholar] [CrossRef]
  36. Lombardo, G.; Arena, G. The use of galantamine bromhydrate (nivaline) in the paralytic sequelae of poliomyelitis, neuraxitis and in muscular dystrophy. Minerva Pediatr. 1962, 30, 724–728. [Google Scholar]
  37. Stoyanov, E.; Vulchanova, S. The clinical application of nivalin as an antidote of curare. Nauchni Tr. Viss. Meditsinski Inst. Sofiia 1963, 42, 45–48. [Google Scholar]
  38. Baraka, A. Reversal of Central Anticholinergic Syndrome by Galanthamine. JAMA 1977, 238, 2293–2294. [Google Scholar] [CrossRef]
  39. Czollner, L.; Frantsits, W.; Küenburg, B.; Hedenig, U.; Fröhlich, J.; Jordis, U. New kilogram-synthesis of the anti-alzheimer drug (−)-galanthamine. Tetrahedron Lett. 1998, 39, 2087–2088. [Google Scholar] [CrossRef]
  40. Albuquerque, E.X.; Santos, M.D.; Alkondon, M.; Pereira, E.F.R.; Maelicke, A. Modulation of Nicotinic Receptor Activity in the Central Nervous System: A Novel Approach to the Treatment of Alzheimer Disease. Alzheimer Dis. Assoc. Disord. 2001, 15, S19–S25. [Google Scholar] [CrossRef]
  41. Dajas-Bailador, F.A.; Heimala, K.; Wonnacott, S. The Allosteric Potentiation of Nicotinic Acetylcholine Receptors by Galantamine Is Transduced into Cellular Responses in Neurons: Ca2+ Signals and Neurotransmitter Release. Mol. Pharmacol. 2003, 64, 1217–1226. [Google Scholar] [CrossRef] [Green Version]
  42. Samochocki, M.; Höffle, A.; Fehrenbacher, A.; Jostock, R.; Ludwig, J.; Christner, C.; Radina, M.; Zerlin, M.; Ullmer, C.; Pereira, E.F.R.; et al. Galantamine Is an Allosterically Potentiating Ligand of Neuronal Nicotinic but Not of Muscarinic Acetylcholine Receptors. J. Pharmacol. Exp. Ther. 2003, 305, 1024–1036. [Google Scholar] [CrossRef] [Green Version]
  43. Schilström, B.; Ivanov, V.B.; Wiker, C.; Svensson, T.H. Galantamine Enhances Dopaminergic Neurotransmission In Vivo Via Allosteric Potentiation of Nicotinic Acetylcholine Receptors. Neuropsychopharmacology 2007, 32, 43–53. [Google Scholar] [CrossRef] [Green Version]
  44. Lilienfeld, S. Galantamine—A Novel Cholinergic Drug with a Unique Dual Mode of Action for the Treatment of Patients with Alzheimer’s Disease. CNS Drug Rev. 2002, 8, 159–176. [Google Scholar] [CrossRef]
  45. Tang, X.C.; Han, Y.F. Pharmacological Profile of Huperzine A, a Novel Acetylcholinesterase Inhibitor from Chinese Herb. CNS Drug Rev. 1999, 5, 281–300. [Google Scholar] [CrossRef] [Green Version]
  46. Wang, Y.; Yue, D.; Tang, X. Anti-Cholinesterase Activity of Huerpzine A. Acta Pharmacol. Sin. 1986, 7, 110–113. [Google Scholar]
  47. Wang, H.; Tang, X. Anticholinesterase Effects of Huperzine A, E2020, and Tacrine in Rats. Zhongguo Yao Li Xue Bao Acta Pharmacol. Sin. 1998, 19, 27–30. [Google Scholar]
  48. Wang, R.; Tang, X.C. Neuroprotective Effects of Huperzine A. Neurosignals 2005, 14, 71–82. [Google Scholar] [CrossRef] [PubMed]
  49. McKinney, M.; Miller, J.H.; Yamada, F.; Tuckmantel, W.; Kozikowski, A.P. Potencies and stereoselectivities of enantiomers of huperzine A for inhibition of rat cortical acetylcholinesterase. Eur. J. Pharmacol. 1991, 203, 303–305. [Google Scholar] [CrossRef]
  50. Dvir, H.; Jiang, H.L.; Wong, D.M.; Harel, M.; Chetrit, M.; He, X.C.; Jin, G.Y.; Yu, G.L.; Tang, X.C.; Silman, I.; et al. X-ray Structures of Torpedo californica Acetylcholinesterase Complexed with (+)-Huperzine A and (−)-Huperzine B: Structural Evidence for an Active Site Rearrangement. Biochemistry 2002, 41, 10810–10818. [Google Scholar] [CrossRef] [PubMed]
  51. Zhang, H.Y.; Liang, Y.Q.; Tang, X.C.; He, X.C.; Bai, D.L. Stereoselectivities of enantiomers of huperzine A in protection against β-amyloid25–35-induced injury in PC12 and NG108-15 cells and cholinesterase inhibition in mice. Neurosci. Lett. 2002, 317, 143–146. [Google Scholar] [CrossRef]
  52. Zhang, H.-Y.; Brimijoin, S.; Tang, X.-C. Apoptosis induced by beta-amyloid25-35 in acetylcholinesterase-overexpressing neuroblastoma cells. Acta Pharmacol. Sin. 2003, 24, 853–858. [Google Scholar]
  53. Yan, X.F.; Lu, W.H.; Lou, W.J.; Tang, X.C. Effects of Huperzine A and B on Skeletal Muscle and the Electroenphalogram. Acta Pharmacol. Sin. 1987, 8, 117–123. [Google Scholar]
  54. Zhang, R.W.; Tang, X.C.; Han, Y.Y.; Sang, G.W.; Zhang, Y.D.; Ma, Y.X.; Zhang, C.L.; Yang, R.M. Drug evaluation of huperzine A in the treatment of senile memory disorders. Acta Pharmacol. Sin. 1991, 12, 250–252. [Google Scholar]
  55. Xu, S.S.; Gao, Z.X.; Weng, Z.; Du, Z.M.; Xu, W.A.; Yang, J.S.; Zhang, M.L.; Tong, Z.H.; Fang, Y.S.; Chai, X.S. Efficacy of Tablet Huperzine A on Memory, Cognition and Behavior in Alzheimer´s Disease. Acta Pharmacol. Sin. 1995, 16, 195–391. [Google Scholar]
  56. Xu, S.S.; Cai, Z.Y.; Qu, Z.W.; Yang, R.M.; Cai, Y.L.; Wang, G.Q.; Su, X.Q.; Zhong, X.S.; Cheng, R.Y.; A Xu, W.; et al. Huperzine-A in capsules and tablets for treating patients with Alzheimer disease. Acta Pharmacol. Sin. 1999, 20, 486–490. [Google Scholar]
  57. Orhan, G.; Orhan, I.; Subutay-Oztekin, N.; Ak, F.; Sener, B. Contemporary Anticholinesterase Pharmaceuticals of Natural Origin and Their Synthetic Analogues for the Treatment of Alzheimers Disease. Recent Patents CNS Drug Discov. 2009, 4, 43–51. [Google Scholar] [CrossRef]
  58. Colovic, M.B.; Krstic, D.Z.; Lazarevic-Pasti, T.D.; Bondzic, A.M.; Vasic, V.M. Acetylcholinesterase Inhibitors: Pharmacology and Toxicology. Curr. Neuropharmacol. 2013, 11, 315–335. [Google Scholar] [CrossRef] [Green Version]
  59. Batiha, G.E.-S.; Alkazmi, L.M.; Nadwa, E.H.; Rashwan, E.K.; Beshbishy, A.M.; Shaheen, H.; Wasef, L. Physostigmine: A Plant Alkaloid Isolated from Physostigma venenosum: A Review on Pharmacokinetics, Pharmacological and Toxicological Activities. J. Drug Deliv. Ther. 2020, 10, 187–190. [Google Scholar] [CrossRef] [Green Version]
  60. Watkins, J.W.; Schwarz, E.S.; Arroyo-Plasencia, A.M.; Mullins, M.E. The Use of Physostigmine by Toxicologists in Anticholinergic Toxicity. J. Med. Toxicol. 2015, 11, 179–184. [Google Scholar] [CrossRef] [Green Version]
  61. Walter, K.; Muller, M.; Barkworth, M.F.; Nieciecki, A.V.; Stanislaus, F. Pharmacokinetics of physostigmine in man following a single application of a transdermal system. Br. J. Clin. Pharmacol. 1995, 39, 59–63. [Google Scholar] [CrossRef] [Green Version]
  62. Allen, N.; Burns, A. The treatment of Alzheimer’s disease. J. Psychopharmacol. 1995, 9, 43–56. [Google Scholar] [CrossRef]
  63. Van Dyck, C.H.; Newhouse, P.; Falk, W.E.; Mattes, J.A. Extended-Release Physostigmine in Alzheimer Disease: A Multicenter, Double-Blind, 12-Week Study With Dose Enrichment. Arch. Gen. Psychiatry 2000, 57, 157–164. [Google Scholar] [CrossRef] [Green Version]
  64. Filho, J.M.J.C.; Birks, J. Physostigmine for dementia due to Alzheimer’s disease. Cochrane Database Syst. Rev. 2001, 2001, CD001499. [Google Scholar] [CrossRef]
  65. Arens, A.M.; Kearney, T. Adverse Effects of Physostigmine. J. Med. Toxicol. 2019, 15, 184–191. [Google Scholar] [CrossRef] [PubMed]
  66. Zhan, Z.-J.; Bian, H.-L.; Wang, J.-W.; Shan, W.-G. Synthesis of physostigmine analogues and evaluation of their anticholinesterase activities. Bioorganic Med. Chem. Lett. 2010, 20, 1532–1534. [Google Scholar] [CrossRef] [PubMed]
  67. Liao, B.; Newmark, H.; Zhou, R. Neuroprotective Effects of Ginseng Total Saponin and Ginsenosides Rb1 and Rg1 on Spinal Cord Neurons In Vitro. Exp. Neurol. 2002, 173, 224–234. [Google Scholar] [CrossRef] [PubMed]
  68. Gillis, C. Panax ginseng pharmacology: A nitric oxide link? Biochem. Pharmacol. 1997, 54, 1–8. [Google Scholar] [CrossRef]
  69. Pena, I.D.; Yoon, S.Y.; Kim, H.J.; Park, S.; Hong, E.Y.; Ryu, J.H.; Park, I.H.; Cheong, J.H. Effects of ginseol k-g3, an Rg3-enriched fraction, on scopolamine-induced memory impairment and learning deficit in mice. J. Ginseng Res. 2014, 38, 1–7. [Google Scholar] [CrossRef] [Green Version]
  70. Rausch, W.-D.; Liu, S.; Gille, G.; Radad, K. Neuroprotective effects of ginsenosides. Acta Neurobiol. Exp. 2006, 66, 369–375. [Google Scholar]
  71. Nah, S.-Y.; Kim, D.-H.; Rhim, H. Ginsenosides: Are Any of them Candidates for Drugs Acting on the Central Nervous System? CNS Drug Rev. 2007, 13, 381–404. [Google Scholar] [CrossRef]
  72. Radad, K.; Moldzio, R.; Rausch, W.-D. Ginsenosides and Their CNS Targets. CNS Neurosci. Ther. 2011, 17, 761–768. [Google Scholar] [CrossRef]
  73. Lee, D.W.; Andersen, J.K. Iron elevations in the aging Parkinsonian brain: A consequence of impaired iron homeostasis? J. Neurochem. 2010, 112, 332–339. [Google Scholar] [CrossRef]
  74. Selkoe, D.J.; Schenk, D. Alzheimer’s Disease: Molecular Understanding Predicts Amyloid-Based Therapeutics. Annu. Rev. Pharmacol. Toxicol. 2003, 43, 545–584. [Google Scholar] [CrossRef]
  75. Dedov, V.N.; Griffiths, F.M.; Garner, B.; Halliday, G.M.; Double, K.L. Lipid content determines aggregation of neuromelanin granules in vitro. J. Neural. Transm. Suppl. 2007, 72, 35–38. [Google Scholar] [CrossRef]
  76. Tanzi, R.E.; Bertram, L. Alzheimer´s Disease: The latest suspect. Nature 2008, 454, 707–708. [Google Scholar] [CrossRef]
  77. Xu, B.-B.; Liu, C.-Q.; Gao, X.; Zhang, W.-Q.; Wang, S.-W.; Cao, Y.-L. Possible mechanisms of the protection of ginsenoside Re against MPTP-induced apoptosis in substantia nigra neurons of Parkinson’s disease mouse model. J. Asian Nat. Prod. Res. 2005, 7, 215–224. [Google Scholar] [CrossRef]
  78. Fang, F.; Chen, X.; Huang, T.; Lue, L.-F.; Luddy, J.S.; Yan, S.S. Multi-faced neuroprotective effects of Ginsenoside Rg1 in an Alzheimer mouse model. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2012, 1822, 286–292. [Google Scholar] [CrossRef] [Green Version]
  79. Hong-Cai, W.; Yu-Meng, J.; Xue, Z.; Ning, W.; Di, G.; Ming, G.; Jing-Yu, Z.; Zi-Chao, Y. Protective Effects of Ginsenoside Rb1 on Aβ Amyloid-Induced Hippocampal Neuronal Injury in Rats. J. Jilin Univ. Med. Ed. 2012, 38, 447–450. [Google Scholar]
  80. Huang, X.; Li, N.; Pu, Y.; Zhang, T.; Wang, B. Neuroprotective Effects of Ginseng Phytochemicals: Recent Perspectives. Molecules 2019, 24, 2939. [Google Scholar] [CrossRef] [Green Version]
  81. Zhao, H.-H.; Di, J.; Liu, W.-S.; Liu, H.-L.; Lai, H.; Lü, Y.-L. Involvement of GSK3 and PP2A in ginsenoside Rb1’s attenuation of aluminum-induced tau hyperphosphorylation. Behav. Brain Res. 2013, 241, 228–234. [Google Scholar] [CrossRef] [PubMed]
  82. Oh, J.; Kim, J.-S. Compound K derived from ginseng: Neuroprotection and cognitive improvement. Food Funct. 2016, 7, 4506–4515. [Google Scholar] [CrossRef] [PubMed]
  83. Oken, B.S.; Storzbach, D.M.; Kaye, J.A. The efficacy of Ginkgo biloba on cognitive function in Alzheimer disease. Arch. Neurol. 1998, 55, 1409–1415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Vellas, B.; Coley, N.; Ousset, P.-J.; Berrut, G.; Dartigues, J.-F.; Dubois, B.; Grandjean, H.; Pasquier, F.; Piette, F.; Robert, P.; et al. Long-term use of standardised ginkgo biloba extract for the prevention of Alzheimer’s disease (GuidAge): A randomised placebo-controlled trial. Lancet Neurol. 2012, 11, 851–859. [Google Scholar] [CrossRef]
  85. Singh, S.K.; Srivastav, S.; Castellani, R.J.; Plascencia-Villa, G.; Perry, G. Neuroprotective and Antioxidant Effect of Ginkgo biloba Extract Against AD and Other Neurological Disorders. Neurotherapeutics 2019, 16, 666–674. [Google Scholar] [CrossRef]
  86. Smith, P.F.; Maclennan, K.; Darlington, C.L. The neuroprotective properties of the Ginkgo biloba leaf: A review of the possible relationship to platelet-activating factor (PAF). J. Ethnopharmacol. 1996, 50, 131–139. [Google Scholar] [CrossRef]
  87. Shi, C.; Liu, J.; Wu, F.; Yew, D.T. Ginkgo biloba Extract in Alzheimer’s Disease: From Action Mechanisms to Medical Practice. Int. J. Mol. Sci. 2010, 11, 107–123. [Google Scholar] [CrossRef]
  88. Augustin, S.; Rimbach, G.; Augustin, K.; Schliebs, R.; Wolffram, S.; Cermak, R. Effect of a short- and long-term treatment with Ginkgo biloba extract on Amyloid Precursor Protein Levels in a transgenic mouse model relevant to Alzheimer’s disease. Arch. Biochem. Biophys. 2009, 481, 177–182. [Google Scholar] [CrossRef]
  89. Löffler, T.; Lee, S.K.; Nöldner, M.; Chatterjee, S.S.; Hoyer, S.; Schliebs, R. Effect of Ginkgo biloba extract (EGb761) on glucose metabolism-related markers in streptozotocin-damaged rat brain. J. Neural Transm. 2001, 108, 1457–1474. [Google Scholar] [CrossRef]
  90. Le Bars, P.L.; Katz, M.M.; Berman, N.; Itil, T.M.; Freedman, A.M.; Schatzberg, A.F. A Placebo-Controlled, Double-blind, Randomized Trial of an Extract of Ginkgo Biloba for Dementia. JAMA 1997, 278, 1327–1332. [Google Scholar] [CrossRef]
  91. Mix, J.A.; Crews, W.D. A double-blind, placebo-controlled, randomized trial of Ginkgo biloba extract EGb 761® in a sample of cognitively intact older adults: Neuropsychological findings. Hum. Psychopharmacol. Clin. Exp. 2002, 17, 267–277. [Google Scholar] [CrossRef]
  92. Kanowski, S.; Herrmann, W.M.; Stephan, K.; Wierich, W.; Hörr, R. Proof of Efficacy of the Ginkgo Biloba Special Extract EGb 761 in Outpatients Suffering from Mild to Moderate Primary Degenerative Dementia of the Alzheimer Type or Multi-infarct Dementia. Pharmacopsychiatry 2007, 29, 47–56. [Google Scholar] [CrossRef]
  93. Napryeyenko, O.; Borzenko, I. Ginkgo biloba Special Extract in Dementia with Neuropsychiatric Features. Arzneimittelforschung 2007, 57, 4–11. [Google Scholar] [CrossRef]
  94. DeKosky, S.T. Ginkgo biloba for Prevention of DementiaA Randomized Controlled Trial. JAMA 2008, 300, 2253–2262. [Google Scholar] [CrossRef]
  95. McCarney, R.; Fisher, P.; Iliffe, S.; van Haselen, R.; Griffin, M.; Van Der Meulen, J.; Warner, J. Ginkgo biloba for mild to moderate dementia in a community setting: A pragmatic, randomised, parallel-group, double-blind, placebo-controlled trial. Int. J. Geriatr. Psychiatry 2008, 23, 1222–1230. [Google Scholar] [CrossRef]
  96. Bhat, A.; Mahalakshmi, A.M.; Ray, B.; Tuladhar, S.; Hediyal, T.A.; Manthiannem, E.; Padamati, J.; Chandra, R.; Chidambaram, S.B.; Sakharkar, M.K. Benefits of curcumin in brain disorders. BioFactors 2019, 45, 666–689. [Google Scholar] [CrossRef]
  97. Khan, M.S.; Ali, T.; Kim, M.W.; Jo, M.H.; Jo, M.G.; Badshah, H.; Kim, M.O. Anthocyanins protect against LPS-induced oxidative stress-mediated neuroinflammation and neurodegeneration in the adult mouse cortex. Neurochem. Int. 2016, 100, 1–10. [Google Scholar] [CrossRef]
  98. Khan, M.S.; Ali, T.; Kim, M.W.; Jo, M.H.; Chung, J.I.; Kim, M.O. Anthocyanins Improve Hippocampus-Dependent Memory Function and Prevent Neurodegeneration via JNK/Akt/GSK3β Signaling in LPS-Treated Adult Mice. Mol. Neurobiol. 2019, 56, 671–687. [Google Scholar] [CrossRef]
  99. Rehman, S.U.; Shah, S.A.; Ali, T.; Chung, J.I.; Kim, M.O. Anthocyanins Reversed D-Galactose-Induced Oxidative Stress and Neuroinflammation Mediated Cognitive Impairment in Adult Rats. Mol. Neurobiol. 2017, 54, 255–271. [Google Scholar] [CrossRef]
  100. Ali, T.; Kim, T.; Rehman, S.U.; Khan, M.S.; Amin, F.U.; Khan, M.; Ikram, M.; Kim, M.O. Natural Dietary Supplementation of Anthocyanins via PI3K/Akt/Nrf2/HO-1 Pathways Mitigate Oxidative Stress, Neurodegeneration, and Memory Impairment in a Mouse Model of Alzheimer’s Disease. Mol. Neurobiol. 2018, 55, 6076–6093. [Google Scholar] [CrossRef]
  101. Qin, L.; Zhang, J.; Qin, M. Protective effect of cyanidin 3-O-glucoside on beta-amyloid peptide-induced cognitive impairment in rats. Neurosci. Lett. 2013, 534, 285–288. [Google Scholar] [CrossRef]
  102. Wang, Q.; Liang, X.; Wang, L.; Lu, X.; Huang, J.; Cao, J.; Li, H.; Gu, D. Effect of omega-3 fatty acids supplementation on endothelial function: A meta-analysis of randomized controlled trials. Atherosclerosis 2012, 221, 536–543. [Google Scholar] [CrossRef]
  103. Bowman, G.L.; Silbert, L.C.; Dodge, H.H.; Lahna, D.; Hagen, K.; Murchison, C.F.; Howieson, D.; Kaye, J.; Quinn, J.F.; Shinto, L. Randomized Trial of Marine n-3 Polyunsaturated Fatty Acids for the Prevention of Cerebral Small Vessel Disease and Inflammation in Aging (PUFA Trial): Rationale, Design and Baseline Results. Nutrients 2019, 11, 735. [Google Scholar] [CrossRef] [Green Version]
  104. Khanna, A.K. Mechanism of the combination immunosuppressive effects of rapamycin with either cyclosporine or tacrolimus. Transplantation 2000, 70, 690–694. [Google Scholar] [CrossRef]
  105. Mohacsi, P.J.; Morris, R.E. Brief treatment with rapamycin in vivo increases responsiveness to alloantigens measured by the mixed lymphocyte response. Immunol. Lett. 1992, 34, 273–278. [Google Scholar] [CrossRef]
  106. Wicker, L.S.; Boltz, R.C.; Matt, V.; Nichols, E.A.; Peterson, L.B.; Sigal, N.H. Suppression of B cell activation by cyclosporin A, FK506 and rapamycin. Eur. J. Immunol. 1990, 20, 2277–2283. [Google Scholar] [CrossRef]
  107. Jiang, J.; Jiang, J.; Zuo, Y.; Gu, Z. Rapamycin protects the mitochondria against oxidative stress and apoptosis in a rat model of Parkinson’s disease. Int. J. Mol. Med. 2013, 31, 825–832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Majumder, S.; Richardson, A.; Strong, R.; Oddo, S. Inducing Autophagy by Rapamycin Before, but Not After, the Formation of Plaques and Tangles Ameliorates Cognitive Deficits. PLoS ONE 2011, 6, e25416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Avrahami, L.; Farfara, D.; Shaham-Kol, M.; Vassar, R.; Frenkel, D.; Eldar-Finkelman, H. Inhibition of Glycogen Synthase Kinase-3 Ameliorates β-Amyloid Pathology and Restores Lysosomal Acidification and Mammalian Target of Rapamycin Activity in the Alzheimer Disease Mouse Model: In vivo and in vitro studies. J. Biol. Chem. 2013, 288, 1295–1306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Caccamo, A.; Majumder, S.; Richardson, A.; Strong, R.; Oddo, S. Molecular Interplay between Mammalian Target of Rapamycin (mTOR), Amyloid-β, and Tau: Effects on Cognivitive Impairments. J. Biol. Chem. 2010, 285, 13107–13120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Siman, R.; Cocca, R.; Dong, Y. The mTOR Inhibitor Rapamycin Mitigates Perforant Pathway Neurodegeneration and Synapse Loss in a Mouse Model of Early-Stage Alzheimer-Type Tauopathy. PLoS ONE 2015, 10, e0142340. [Google Scholar] [CrossRef]
  112. Lin, A.-L.; Jahrling, J.B.; Zhang, W.; DeRosa, N.; Bakshi, V.; Romero, P.; Galvan, V.; Richardson, A. Rapamycin rescues vascular, metabolic and learning deficits in apolipoprotein E4 transgenic mice with pre-symptomatic Alzheimer’s disease. J. Cereb. Blood Flow Metab. 2017, 37, 217–226. [Google Scholar] [CrossRef] [Green Version]
  113. Lin, A.-L.; Zheng, W.; Halloran, J.J.; Burbank, R.R.; A Hussong, S.; Hart, M.J.; A Javors, M.; Shih, Y.-Y.I.; Muir, E.R.; Fonseca, R.S.; et al. Chronic Rapamycin Restores Brain Vascular Integrity and Function Through NO Synthase Activation and Improves Memory in Symptomatic Mice Modeling Alzheimer’s Disease. J. Cereb. Blood Flow Metab. 2013, 33, 1412–1421. [Google Scholar] [CrossRef]
  114. Walter, L.; Franklin, A.; Witting, A.; Wade, C.; Xie, Y.; Kunos, G.; Mackie, K.; Stella, N. Nonpsychotropic Cannabinoid Receptors Regulate Microglial Cell Migration. J. Neurosci. 2003, 23, 1398–1405. [Google Scholar] [CrossRef] [Green Version]
  115. Hamelink, C.; Hampson, A.; Wink, D.A.; Eiden, L.E.; Eskay, R.L. Comparison of cannabidiol, antioxidants, and diuretics in reversing binge ethanol-induced neurotoxicity. J. Pharmacol. Exp. Ther. 2005, 314, 780–788. [Google Scholar] [CrossRef] [Green Version]
  116. Esposito, G.; De Filippis, D.; Carnuccio, R.; Izzo, A.; Iuvone, T. The marijuana component cannabidiol inhibits β-amyloid-induced tau protein hyperphosphorylation through Wnt/β-catenin pathway rescue in PC12 cells. Klin. Wochenschr. 2006, 84, 253–258. [Google Scholar] [CrossRef]
  117. Esposito, G.; De Filippis, D.; Maiuri, M.C.; De Stefano, D.; Carnuccio, R.; Iuvone, T. Cannabidiol inhibits inducible nitric oxide synthase protein expression and nitric oxide production in β-amyloid stimulated PC12 neurons through p38 MAP kinase and NF-κB involvement. Neurosci. Lett. 2006, 399, 91–95. [Google Scholar] [CrossRef]
  118. Martín-Moreno, A.M.; Reigada, D.; Ramírez, B.G.; Mechoulam, R.; Innamorato, N.; Cuadrado, A.; de Ceballos, M.L. Cannabidiol and Other Cannabinoids Reduce Microglial Activation In Vitro and In Vivo: Relevance to Alzheimer’s Disease. Mol. Pharmacol. 2011, 79, 964–973. [Google Scholar] [CrossRef] [Green Version]
  119. Mukhopadhyay, P.; Rajesh, M.; Horváth, B.; Bátkai, S.; Park, O.; Tanchian, G.; Gao, R.Y.; Patel, V.; Wink, D.A.; Liaudet, L.; et al. Cannabidiol protects against hepatic ischemia/reperfusion injury by attenuating inflammatory signaling and response, oxidative/nitrative stress, and cell death. Free Radic. Biol. Med. 2011, 50, 1368–1381. [Google Scholar] [CrossRef] [Green Version]
  120. Esposito, G.; Scuderi, C.; Savani, C.; Steardo, L., Jr.; De Filippis, D.; Cottone, P.; Iuvone, T.; Cuomo, V.; Steardo, L. Cannabidiol in vivo blunts β-amyloid induced neuroinflammation by suppressing IL-1β and iNOS expression: CBD Blunts Aβ Induced Neuroinflammation In Vivo. Br. J. Pharmacol. 2007, 151, 1272–1279. [Google Scholar] [CrossRef] [Green Version]
  121. Esposito, G.; Scuderi, C.; Valenza, M.; Togna, G.I.; Latina, V.; De Filippis, D.; Cipriano, M.; Carratù, M.R.; Iuvone, T.; Steardo, L. Cannabidiol Reduces Aβ-Induced Neuroinflammation and Promotes Hippocampal Neurogenesis through PPARγ Involvement. PLoS ONE 2011, 6, e28668. [Google Scholar] [CrossRef] [PubMed]
  122. Wang, X.; Song, R.; Lu, W.; Liu, Z.; Wang, L.; Zhu, X.; Liu, Y.; Sun, Z.; Li, J.; Li, X. YXQN Reduces Alzheimer’s Disease-Like Pathology and Cognitive Decline in APPswePS1dE9 Transgenic Mice. Front. Aging Neurosci. 2017, 9, 157. [Google Scholar] [CrossRef] [PubMed]
  123. Gao, Y.; Zhang, L.; Jiao, W. Marine glycan-derived therapeutics in China. Prog. Mol. Biol. Transl. Sci. 2019, 163, 113–134. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, X.; Sun, G.; Feng, T.; Zhang, J.; Huang, X.; Wang, T.; Xie, Z.; Chu, X.; Yang, J.; Wang, H.; et al. Sodium oligomannate therapeutically remodels gut microbiota and suppresses gut bacterial amino acids-shaped neuroinflammation to inhibit Alzheimer’s disease progression. Cell Res. 2019, 29, 787–803. [Google Scholar] [CrossRef]
  125. Syed, Y.Y. Sodium Oligomannate: First Approval. Drugs 2020, 80, 441–444. [Google Scholar] [CrossRef]
  126. Xiao, S.; Chan, P.; Wang, T.; Hong, Z.; Wang, S.; Kuang, W.; He, J.; Pan, X.; Zhou, Y.; Ji, Y.; et al. A 36-week multicenter, randomized, double-blind, placebo-controlled, parallel-group, phase 3 clinical trial of sodium oligomannate for mild-to-moderate Alzheimer’s dementia. Alzheimer’s Res. Ther. 2021, 13, 62. [Google Scholar] [CrossRef]
  127. Witvrouw, M.; De Clercq, E. Sulfated Polysaccharides Extracted from Sea Algae as Potential Antiviral Drugs. Gen. Pharmacol. Vasc. Syst. 1997, 29, 497–511. [Google Scholar] [CrossRef]
  128. Eppley, R.W. Sodium exclusion and potassium retention by the red marine alga, porphyra perforata. J. Gen. Physiol. 1958, 41, 901–911. [Google Scholar] [CrossRef] [Green Version]
  129. Zhang, Q.; Yu, P.; Li, Z.; Zhang, H.; Xu, Z.; Li, P. Antioxidant activities of sulfated polysaccharide fractions from Porphyra haitanesis. J. Appl. Phycol. 2003, 15, 305–310. [Google Scholar] [CrossRef]
  130. Isaka, S.; Cho, K.; Nakazono, S.; Abu, R.; Ueno, M.; Kim, D.; Oda, T. Antioxidant and anti-inflammatory activities of porphyran isolated from discolored nori (Porphyra yezoensis). Int. J. Biol. Macromol. 2015, 74, 68–75. [Google Scholar] [CrossRef]
  131. Liu, X.-Y.; Liu, D.; Lin, G.-P.; Wu, Y.-J.; Gao, L.-Y.; Ai, C.; Huang, Y.-F.; Wang, M.-F.; El-Seedi, H.R.; Chen, X.-H.; et al. Anti-ageing and antioxidant effects of sulfate oligosaccharides from green algae Ulva lactuca and Enteromorpha prolifera in SAMP8 mice. Int. J. Biol. Macromol. 2019, 139, 342–351. [Google Scholar] [CrossRef]
  132. Kim, M.; Li, Y.-X.; Dewapriya, P.; Ryu, B.; Kim, S.-K. Floridoside suppresses pro-inflammatory responses by blocking MAPK signaling in activated microglia. BMB Rep. 2013, 46, 398–403. [Google Scholar] [CrossRef]
  133. Zhou, R.; Shi, X.-Y.; Bi, D.-C.; Fang, W.-S.; Wei, G.-B.; Xu, X. Alginate-Derived Oligosaccharide Inhibits Neuroinflammation and Promotes Microglial Phagocytosis of β-Amyloid. Mar. Drugs 2015, 13, 5828–5846. [Google Scholar] [CrossRef] [Green Version]
  134. Cui, Y.-Q.; Jia, Y.-J.; Zhang, T.; Zhang, Q.-B.; Wang, X.-M. Fucoidan Protects against Lipopolysaccharide-Induced Rat Neuronal Damage and Inhibits the Production of Proinflammatory Mediators in Primary Microglia. CNS Neurosci. Ther. 2012, 18, 827–833. [Google Scholar] [CrossRef]
  135. Park, H.Y.; Han, M.H.; Park, C.; Jin, C.-Y.; Kim, G.-Y.; Choi, I.-W.; Kim, N.D.; Nam, T.-J.; Kwon, T.K.; Choi, Y.H. Anti-inflammatory effects of fucoidan through inhibition of NF-κB, MAPK and Akt activation in lipopolysaccharide-induced BV2 microglia cells. Food Chem. Toxicol. 2011, 49, 1745–1752. [Google Scholar] [CrossRef]
  136. Bi, D.; Lai, Q.; Han, Q.; Cai, N.; He, H.; Fang, W.; Yi, J.; Li, X.; Xu, H.; Li, X.; et al. Seleno-polymannuronate attenuates neuroinflammation by suppressing microglial and astrocytic activation. J. Funct. Foods 2018, 51, 113–120. [Google Scholar] [CrossRef]
  137. Yao, Z.-A.; Xu, L.; Wu, H.-G. Immunomodulatory Function of κ-Carrageenan Oligosaccharides Acting on LPS-Activated Microglial Cells. Neurochem. Res. 2014, 39, 333–343. [Google Scholar] [CrossRef]
  138. Schilling, S.; Zeitschel, U.; Hoffmann, T.; Heiser, U.; Francke, M.; Kehlen, A.; Holzer, M.; Hutter-Paier, B.; Prokesch, M.; Windisch, M.; et al. Glutaminyl cyclase inhibition attenuates pyroglutamate Aβ and Alzheimer’s disease–like pathology. Nat. Med. 2008, 14, 1106–1111. [Google Scholar] [CrossRef]
  139. Morawski, M.; Schilling, S.; Kreuzberger, M.; Waniek, A.; Jäger, C.; Koch, B.; Cynis, H.; Kehlen, A.; Arendt, T.; Hartlage-Rübsamen, M.; et al. Glutaminyl Cyclase in Human Cortex: Correlation with (pGlu)-Amyloid-β Load and Cognitive Decline in Alzheimer’s Disease. J. Alzheimer’s Dis. 2014, 39, 385–400. [Google Scholar] [CrossRef] [Green Version]
  140. Bayer, T.A. Pyroglutamate Aβ cascade as drug target in Alzheimer’s disease. Mol. Psychiatry 2021, 1–6. [Google Scholar] [CrossRef]
  141. Hellvard, A.; Maresz, K.; Schilling, S.; Graubner, S.; Heiser, U.; Jonsson, R.; Cynis, H.; DeMuth, H.-U.; Potempa, J.; Mydel, P. Glutaminyl Cyclases as Novel Targets for the Treatment of Septic Arthritis. J. Infect. Dis. 2013, 207, 768–777. [Google Scholar] [CrossRef] [Green Version]
  142. Scheltens, P.; Hallikainen, M.; Grimmer, T.; Duning, T.; A Gouw, A.; E Teunissen, C.; Wink, A.M.; Maruff, P.; Harrison, J.; Van Baal, C.M.; et al. Safety, tolerability and efficacy of the glutaminyl cyclase inhibitor PQ912 in Alzheimer’s disease: Results of a randomized, double-blind, placebo-controlled phase 2a study. Alzheimer’s Res. Ther. 2018, 10, 107. [Google Scholar] [CrossRef]
  143. Hielscher-Michael, S.; Griehl, C.; Buchholz, M.; DeMuth, H.-U.; Arnold, N.; Wessjohann, L.A. Natural Products from Microalgae with Potential against Alzheimer’s Disease: Sulfolipids Are Potent Glutaminyl Cyclase Inhibitors. Mar. Drugs 2016, 14, 203. [Google Scholar] [CrossRef]
  144. Al-Awadhi, F.; Ratnayake, R.; Paul, V.J.; Luesch, H. Tasiamide F, a potent inhibitor of cathepsins D and E from a marine cyanobacterium. Bioorganic Med. Chem. 2016, 24, 3276–3282. [Google Scholar] [CrossRef] [Green Version]
  145. Liu, Y.; Zhang, W.; Li, L.; Salvador, L.A.; Chen, T.; Chen, W.; Felsenstein, K.M.; Ladd, T.B.; Price, A.R.; Golde, T.E.; et al. Cyanobacterial Peptides as a Prototype for the Design of Potent β-Secretase Inhibitors and the Development of Selective Chemical Probes for Other Aspartic Proteases. J. Med. Chem. 2012, 55, 10749–10765. [Google Scholar] [CrossRef]
  146. Rafiquzzaman, S.; Kim, E.Y.; Lee, J.M.; Mohibbullah; Alam, B.; Moon, I.S.; Kim, J.-M.; Kong, I.-S. Anti-Alzheimers and anti-inflammatory activities of a glycoprotein purified from the edible brown alga Undaria pinnatifida. Food Res. Int. 2015, 77, 118–124. [Google Scholar] [CrossRef]
  147. Lee, S.; Youn, K.; Kim, D.H.; Ahn, M.-R.; Yoon, E.; Kim, O.-Y.; Jun, M. Anti-Neuroinflammatory Property of Phlorotannins from Ecklonia cava on Aβ25-35-Induced Damage in PC12 Cells. Mar. Drugs 2018, 17, 7. [Google Scholar] [CrossRef] [Green Version]
  148. Jung, W.-K.; Heo, S.-J.; Jeon, Y.-J.; Lee, C.-M.; Park, Y.-M.; Byun, H.-G.; Choi, Y.H.; Park, S.-G.; Choi, I.-W. Inhibitory Effects and Molecular Mechanism of Dieckol Isolated from Marine Brown Alga on COX-2 and iNOS in Microglial Cells. J. Agric. Food Chem. 2009, 57, 4439–4446. [Google Scholar] [CrossRef] [PubMed]
  149. Zou, Y.; Qian, Z.-J.; Li, Y.; Kim, M.-M.; Lee, S.-H.; Kim, S.-K. Antioxidant Effects of Phlorotannins Isolated from Ishige okamurae in Free Radical Mediated Oxidative Systems. J. Agric. Food Chem. 2008, 56, 7001–7009. [Google Scholar] [CrossRef] [PubMed]
  150. Heo, S.-J.; Kim, J.-P.; Jung, W.-K.; Lee, N.-H.; Kang, H.-S.; Jun, E.-M.; Park, S.-H.; Kang, S.-M.; Lee, Y.-J.; Park, P.-J.; et al. Identification of Chemical Structure and Free Radical Scavenging Activity of Diphlorehtohydroxycarmalol Isolated from a Brown Alga, Ishige Okamurae. J. Microbiol. Biotechnol. 2008, 18, 676–681. [Google Scholar] [PubMed]
  151. Yu, D.-K.; Lee, B.; Kwon, M.; Yoon, N.; Shin, T.; Kim, N.-G.; Choi, J.-S.; Kim, H.-R. Phlorofucofuroeckol B suppresses inflammatory responses by down-regulating nuclear factor κB activation via Akt, ERK, and JNK in LPS-stimulated microglial cells. Int. Immunopharmacol. 2015, 28, 1068–1075. [Google Scholar] [CrossRef]
  152. Jung, H.A.; Jin, S.E.; Ahn, B.R.; Lee, C.M.; Choi, J.S. Anti-inflammatory activity of edible brown alga Eisenia bicyclis and its constituents fucosterol and phlorotannins in LPS-stimulated RAW264.7 macrophages. Food Chem. Toxicol. 2013, 59, 199–206. [Google Scholar] [CrossRef]
  153. Kim, A.-R.; Lee, M.-S.; Choi, J.-W.; Utsuki, T.; Kim, J.-I.; Jang, B.-C.; Kim, H.-R. Phlorofucofuroeckol A Suppresses Expression of Inducible Nitric Oxide Synthase, Cyclooxygenase-2, and Pro-inflammatory Cytokines via Inhibition of Nuclear Factor-κB, c-Jun NH2-Terminal Kinases, and Akt in Microglial Cells. Inflammation 2013, 36, 259–271. [Google Scholar] [CrossRef]
  154. Yang, Y.-I.; Jung, S.-H.; Lee, K.-T.; Choi, J.-H. 8,8′-Bieckol, isolated from edible brown algae, exerts its anti-inflammatory effects through inhibition of NF-κB signaling and ROS production in LPS-stimulated macrophages. Int. Immunopharmacol. 2014, 23, 460–468. [Google Scholar] [CrossRef]
  155. Kim, A.-R.; Lee, B.; Joung, E.-J.; Gwon, W.-G.; Utsuki, T.; Kim, N.-G.; Kim, H.-R. 6,6′-Bieckol suppresses inflammatory responses by down-regulating nuclear factor-κB activation via Akt, JNK, and p38 MAPK in LPS-stimulated microglial cells. Immunopharmacol. Immunotoxicol. 2016, 38, 244–252. [Google Scholar] [CrossRef]
  156. Kannan, R.R.; Aderogba, M.A.; Ndhlala, A.R.; Stirk, W.A.; Van Staden, J. Acetylcholinesterase inhibitory activity of phlorotannins isolated from the brown alga, Ecklonia maxima (Osbeck) Papenfuss. Food Res. Int. 2013, 54, 1250–1254. [Google Scholar] [CrossRef]
  157. Lee, B.H.; Choi, B.W.; Lee, S.Y. Isolation of 6,6’-Bieckol from Grateloupia elliptica and its Antioxidative and Anti-Cholinesterase Activity. Ocean Polar Res. 2017, 39, 45–49. [Google Scholar] [CrossRef]
  158. Yang, E.-J.; Ham, Y.M.; Yang, K.-W.; Lee, N.H.; Hyun, C.-G. Sargachromenol fromSargassum micracanthumInhibits the Lipopolysaccharide-Induced Production of Inflammatory Mediators in RAW 264.7 Macrophages. Sci. World J. 2013, 2013, 712303. [Google Scholar] [CrossRef] [Green Version]
  159. Kang, G.-J.; Han, S.-C.; Yoon, W.-J.; Koh, Y.-S.; Hyun, J.-W.; Kang, H.-K.; Cho, J.Y.; Yoo, E.-S. Sargaquinoic acid isolated fromSargassum siliquastruminhibits lipopolysaccharide-induced nitric oxide production in macrophagesviamodulation of nuclear factor-κB and c-JunN-terminal kinase pathways. Immunopharmacol. Immunotoxicol. 2013, 35, 80–87. [Google Scholar] [CrossRef]
  160. Myung, C.-S.; Shin, H.-C.; Bao, H.Y.; Yeo, S.J.; Lee, B.H.; Kang, J.S. Improvement of memory by dieckol and phlorofucofuroeckol in ethanol-treated mice: Possible involvement of the inhibition of acetylcholinesterase. Arch. Pharm. Res. 2005, 28, 691–698. [Google Scholar] [CrossRef]
  161. Choi, B.W.; Ryu, G.; Park, S.H.; Kim, E.S.; Shin, J.; Roh, S.S.; Shin, H.C.; Lee, B.H. Anticholinesterase activity of plastoquinones fromSargassum sagamianum: Lead compounds for alzheimer’s disease therapy. Phytother. Res. 2007, 21, 423–426. [Google Scholar] [CrossRef]
  162. Seong, S.H.; Ali, M.Y.; Kim, H.-R.; Jung, H.A.; Choi, J.S. BACE1 inhibitory activity and molecular docking analysis of meroterpenoids from Sargassum serratifolium. Bioorg. Med. Chem. 2017, 25, 3964–3970. [Google Scholar] [CrossRef]
  163. Seong, S.H.; Paudel, P.; Jung, H.A.; Choi, J.S. Identifying Phlorofucofuroeckol-A as a Dual Inhibitor of Amyloid-β25-35 Self-Aggregation and Insulin Glycation: Elucidation of the Molecular Mechanism of Action. Mar. Drugs 2019, 17, 600. [Google Scholar] [CrossRef] [Green Version]
  164. Gonçalves, K.G.; da Silva, L.L.; Soares, A.R.; Romeiro, N.C. Acetylcholinesterase as a target of halogenated marine natural products from Laurencia dendroidea. Algal Res. 2020, 52, 102130. [Google Scholar] [CrossRef]
  165. Paudel, P.; Seong, S.H.; Zhou, Y.; Park, H.J.; Jung, H.A.; Choi, J.S. Anti-Alzheimer’s Disease Activity of Bromophenols from a Red Alga, Symphyocladia latiuscula (Harvey) Yamada. ACS Omega 2019, 4, 12259–12270. [Google Scholar] [CrossRef] [Green Version]
  166. Wang, J.; Zheng, J.; Huang, C.; Zhao, J.; Lin, J.; Zhou, X.; Naman, C.B.; Wang, N.; Gerwick, W.H.; Wang, Q.; et al. Eckmaxol, a Phlorotannin Extracted from Ecklonia maxima, Produces Anti-β-amyloid Oligomer Neuroprotective Effects Possibly via Directly Acting on Glycogen Synthase Kinase 3β. ACS Chem. Neurosci. 2018, 9, 1349–1356. [Google Scholar] [CrossRef]
  167. Lee, J.K.; Byun, H.-G. A novel BACE inhibitor isolated from Eisenia bicyclis exhibits neuroprotective activity against β-amyloid toxicity. Fish. Aquat. Sci. 2018, 21, 38. [Google Scholar] [CrossRef] [Green Version]
  168. Lee, S.; Lee, Y.S.; Jung, S.H.; Kang, S.S.; Shin, K.H. Anti-oxidant activities of fucosterol from the marine algae Pelvetia siliquosa. Arch. Pharm. Res. 2003, 26, 719–722. [Google Scholar] [CrossRef] [PubMed]
  169. Choi, J.S.; Han, Y.R.; Byeon, J.S.; Choung, S.-Y.; Sohn, H.S.; Jung, H.A. Protective effect of fucosterol isolated from the edible brown algae, Ecklonia stolonifera and Eisenia bicyclis, on tert-butyl hydroperoxide- and tacrine-induced HepG2 cell injury. J. Pharm. Pharmacol. 2015, 67, 1170–1178. [Google Scholar] [CrossRef] [PubMed]
  170. Fernando, I.P.S.; Jayawardena, T.U.; Kim, H.-S.; Lee, W.W.; Vaas, A.P.J.P.; De Silva, H.I.C.; Abayaweera, G.S.; Nanayakkara, C.M.; Abeytunga, D.T.U.; Lee, D.-S.; et al. Beijing urban particulate matter-induced injury and inflammation in human lung epithelial cells and the protective effects of fucosterol from Sargassum binderi (Sonder ex J. Agardh). Environ. Res. 2019, 172, 150–158. [Google Scholar] [CrossRef] [PubMed]
  171. Yoo, M.-S.; Shin, J.-S.; Choi, H.-E.; Cho, Y.-W.; Bang, M.-H.; Baek, N.-I.; Lee, K.-T. Fucosterol isolated from Undaria pinnatifida inhibits lipopolysaccharide-induced production of nitric oxide and pro-inflammatory cytokines via the inactivation of nuclear factor-κB and p38 mitogen-activated protein kinase in RAW264.7 macrophages. Food Chem. 2012, 135, 967–975. [Google Scholar] [CrossRef]
  172. Wong, C.H.; Gan, S.Y.; Tan, S.C.; Gany, S.A.; Ying, T.; Gray, A.I.; Igoli, J.; Chan, E.W.L.; Phang, S.M. Fucosterol inhibits the cholinesterase activities and reduces the release of pro-inflammatory mediators in lipopolysaccharide and amyloid-induced microglial cells. J. Appl. Phycol. 2018, 30, 3261–3270. [Google Scholar] [CrossRef]
  173. Sun, Z.; Mohamed, M.A.A.; Park, S.Y.; Yi, T.H. Fucosterol protects cobalt chloride induced inflammation by the inhibition of hypoxia-inducible factor through PI3K/Akt pathway. Int. Immunopharmacol. 2015, 29, 642–647. [Google Scholar] [CrossRef]
  174. Yoon, N.Y.; Chung, H.Y.; Kim, H.R.; Choi, J.S. Acetyl- and butyrylcholinesterase inhibitory activities of sterols and phlorotannins from Ecklonia stolonifera. Fish. Sci. 2008, 74, 200–207. [Google Scholar] [CrossRef]
  175. Castro-Silva, E.S.; Bello, M.; Hernández-Rodríguez, M.; Correa-Basurto, J.; Murillo-Álvarez, J.I.; Rosales-Hernández, M.C.; Muñoz-Ochoa, M. In Vitro and in silico evaluation of fucosterol from Sargassum horridum as potential human acetylcholinesterase inhibitor. J. Biomol. Struct. Dyn. 2019, 37, 3259–3268. [Google Scholar] [CrossRef]
  176. Jung, H.A.; Ali, M.Y.; Choi, R.J.; Jeong, H.O.; Chung, H.Y.; Choi, J.S. Kinetics and molecular docking studies of fucosterol and fucoxanthin, BACE1 inhibitors from brown algae Undaria pinnatifida and Ecklonia stolonifera. Food Chem. Toxicol. 2016, 89, 104–111. [Google Scholar] [CrossRef]
  177. Pangestuti, R.; Vo, T.-S.; Ngo, D.-H.; Kim, S.-K. Fucoxanthin Ameliorates Inflammation and Oxidative Reponses in Microglia. J. Agric. Food Chem. 2013, 61, 3876–3883. [Google Scholar] [CrossRef]
  178. Heo, S.-J.; Ko, S.-C.; Kang, S.-M.; Kang, H.-S.; Kim, J.-P.; Kim, S.-H.; Lee, K.-W.; Cho, M.-G.; Jeon, Y.-J. Cytoprotective effect of fucoxanthin isolated from brown algae Sargassum siliquastrum against H2O2-induced cell damage. Eur. Food Res. Technol. 2008, 228, 145–151. [Google Scholar] [CrossRef]
  179. Zhao, D.; Kwon, S.-H.; Chun, Y.S.; Gu, M.-Y.; Yang, H.O. Anti-Neuroinflammatory Effects of Fucoxanthin via Inhibition of Akt/NF-κB and MAPKs/AP-1 Pathways and Activation of PKA/CREB Pathway in Lipopolysaccharide-Activated BV-2 Microglial Cells. Neurochem. Res. 2017, 42, 667–677. [Google Scholar] [CrossRef]
  180. Kawee-Ai, A.; Kuntiya, A.; Kim, S.M. Anticholinesterase and Antioxidant Activities of Fucoxanthin Purified from the Microalga Phaeodactylum Tricornutum. Nat. Prod. Commun. 2013, 8, 1381–1386. [Google Scholar] [CrossRef] [Green Version]
  181. Jiang, X.; Chen, L.; Shen, L.; Chen, Z.; Xu, L.; Zhang, J.; Yu, X. Trans-astaxanthin attenuates lipopolysaccharide-induced neuroinflammation and depressive-like behavior in mice. Brain Res. 2016, 1649, 30–37. [Google Scholar] [CrossRef] [PubMed]
  182. Shanmuganathan, B.; Malar, D.S.; Sathya, S.; Devi, K.P. Antiaggregation Potential of Padina gymnospora against the Toxic Alzheimer’s Beta-Amyloid Peptide 25–35 and Cholinesterase Inhibitory Property of Its Bioactive Compounds. PLoS ONE 2015, 10, e0141708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Lin, J.; Huang, L.; Yu, J.; Xiang, S.; Wang, J.; Zhang, J.; Yan, X.; Cui, W.; He, S.; Wang, Q. Fucoxanthin, A Marine Carotenoid, Reverses Scopolamine-Induced Cognitive Impairments in Mice and Inhibits Acetylcholinesterase In Vitro. Mar. Drugs 2016, 14, 67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Xiang, S.; Liu, F.; Lin, J.; Chen, H.; Huang, C.; Chen, L.; Zhou, Y.; Ye, L.; Zhang, K.; Jin, J.; et al. Fucoxanthin Inhibits β-Amyloid Assembly and Attenuates β-Amyloid Oligomer-Induced Cognitive Impairments. J. Agric. Food Chem. 2017, 65, 4092–4102. [Google Scholar] [CrossRef]
  185. Alghazwi, M.; Smid, S.; Musgrave, I.; Zhang, W. In Vitro studies of the neuroprotective activities of astaxanthin and fucoxanthin against amyloid beta (Aβ1-42) toxicity and aggregation. Neurochem. Int. 2019, 124, 215–224. [Google Scholar] [CrossRef] [Green Version]
  186. Shanmuganathan, B.; Suryanarayanan, V.; Sathya, S.; Narenkumar, M.; Singh, S.K.; Ruckmani, K.; Devi, K.P. Anti-amyloidogenic and anti-apoptotic effect of α-bisabolol against Aβ induced neurotoxicity in PC12 cells. Eur. J. Med. Chem. 2018, 143, 1196–1207. [Google Scholar] [CrossRef]
  187. Zhang, X.-S.; Zhang, X.; Wu, Q.; Li, W.; Wang, C.-X.; Xie, G.-B.; Zhou, X.-M.; Shi, J.-X.; Zhou, M.-L. Astaxanthin offers neuroprotection and reduces neuroinflammation in experimental subarachnoid hemorrhage. J. Surg. Res. 2014, 192, 206–213. [Google Scholar] [CrossRef]
  188. Soininen, H.; Solomon, A.; Visser, P.J.; Hendrix, S.B.; Blennow, K.; Kivipelto, M.; Hartmann, T.; the LipiDiDiet clinical study group. 36-month LipiDiDiet multinutrient clinical trial in prodromal Alzheimer’s disease. Alzheimer’s Dement. 2021, 17, 29–40. [Google Scholar] [CrossRef]
Table 1. Natural agents in Clinical trials of Alzheimer’s disease drug development (US National Library of Medicine. Available online: https://clinicaltrials.gov, accessed from September 2021 to November 2021.
Table 1. Natural agents in Clinical trials of Alzheimer’s disease drug development (US National Library of Medicine. Available online: https://clinicaltrials.gov, accessed from September 2021 to November 2021.
AgentMechanism of ActionTherapeutic PurposeTrial Identifier and StatusPhase
Huperzine AAChE inhibitor, inhibition of Aβimprove memoryNot yet recruiting
NCT02931136
IV
Sodium oligomannate
(GV-971)
neuroinflammation modulators, microbiome modulators, amyloid beta-protein inhibitors;
reconditioning the dysbiosis of gut microbiota, preventing peripheral immune cells from invading the brain, inhibiting the inflammatory response in the brain targeting protein folding errors in the brain tissue
improve the cognitive function of patients with mild to moderate ADRecruiting
NCT05058040
IV
Sodium oligomannte capsules
(GV-971)
neuroinflammation modulators, microbiome modulators, amyloid beta-protein inhibitors;
reconditioning the dysbiosis of gut microbiota, preventing peripheral immune cells from invading the brain, inhibiting the inflammatory response in the brain targeting protein folding errors in the brain tissue
improve the cognitive function of patients with mild to moderate ADRecruiting
NCT05181475
IV
Ginkgo bilobametabolism and bioenergetics; plant extract with antioxidant propertiesImprove brain blood flow and mitochondrial function (cognitive enhancer)Recruiting
NCT03090516
III
Sodium oligomannate
(GV-971)
reconditioning the dysbiosis of gut microbiota, preventing peripheral immune cells from invading the brain, inhibiting the inflammatory response in the brain targeting protein folding errors in the brain tissueimprove the cognitive function of patients with mild to moderate AD; evaluate safety, tolerability and efficacy of GV-971Recruiting
NCT04520412
III
Curcumin + aerobic yogaherb with antioxidant and anti-inflammatory propertiesdecrease inflammation and oxidation related neurotoxicityactive, not recruiting
NCT01811381
II
Elderberry Juicerich in anthocyanins, has anti-inflammatory and antioxidant activityimprove mitochondrial functioncompleted
NCT02414607
II
Grape powderantioxidant, anti-inflammatory and anticarcinogenicimproves cognitive performance preservation of metabolism in brain regions important to cognitive functionrecruiting
NCT03361410
II
Icosapent ethyl (IPE)synaptic plasticity, neuroprotection; purified from of the omega-3 fatty acid EPAimprove synaptic function; reduce inflammationrecruiting
NCT02719327
II
Meganatrual-Az Grapeseed Extractpolyphenolic extract with antioxidant propertiesanti-oligomerization agent; prevents aggregation of amyloid and taurecruiting
NCT02033941
II
Omega-3 PUFAfish oil concentrate standardized to long chain in n-3 PUFA contentreduces inflammation and glial activation; enhances amyloid removal; protect small blood vesselsactive, not recruiting
NCT01953705
II
Rapamycinanti-inflammatory, antineoplastic; macrolide compound from Streptomyces hygroscopicusselectively blocks the transcriptional activation of cytokinesrecruiting
NCT04629495
II
Rifaximininflammation, infection and immunity; antibioticreduce proinflammatory cytokines secreted by harmful gut bacteriacompleted
NCT03856359
II
Tacrolimustau proteins; macrolide from culture broth of a strain of Streptomyces tsukubaensisreduce pathological changes of tau proteinswithdrawn
NCT04263519
II
THC-free CBD Oilanti-oxidant and anti-inflammatory; cannabinoidsbehavioural and psychological symptoms of dementia (BPSD) decrease with use of cannabinoidsrecruiting
NCT04436081
II
VGH-AD1undisclosed; traditional Chinese herbal medicineundisclosed (cognitive enhancer)not yet recruiting
NCT04249869
II
Yangxue Qingnao pillsblood circulation; traditional Chinese medicine, composed of Angelicae Sinensis Radix, Chuanxiong Rhizoma, Paeoniae Radix Alba, Rhemannia glutinosa, Uncaria macrophylla Wall, Caulis spatholobi, Spica Prunellae, Catsia tora Linn, Mater Margarita, Corydalis ambigua and Asarum sieboldiiimprove cerebral blood flow and brain nourishmentnot yet recruiting
NCT04780399
II
BDPP (bioactive dietary polyphenol preparation)metabolism and bioenergetics, amyloid; combination of grape seed polyphenolic extract and resveratrolprevents amyloid and tau aggregationrecruiting
NCT02502253
I
Pomace olive oilprevent inflammation; lipophilic minor componentsconsumption of olive oil reduces activation of microglia by TRL (triglyceride-rich lipoproteins)completed
NCT04559828
not applicable
Extra virgin olive oil “Coratina”anti-amyloid; biophenolimprove cerebral performancenot yet recruiting
NCT04229186
not applicable
Table 2. Chemical structures and characteristics of esterase inhibitors.
Table 2. Chemical structures and characteristics of esterase inhibitors.
NameStructureSourceCharacteristicsRef.
galantamine Biomolecules 12 00694 i001Galanthus nivalisreversible, competitive AChE inhibitor, allosteric modulator of nicotinic acetylcholine receptors, modulates α4β2 and α7 nicotinic receptors[40,41,42,43]
huperzine A Biomolecules 12 00694 i002Huperzia serrataspecific and reversible AChE inhibitor, protects cells against hydrogen peroxide, β-amyloid toxicity, glutamate, ischemia and staurosporine-induced cytotoxicity and apoptosis[45,46,47,48,51]
physostigmine Biomolecules 12 00694 i003Physostigma venenosum, Streptomyces pseudogriseolusAChE inhibitor[57]
tolserine Biomolecules 12 00694 i004Physostigmine derivativeAChE inhibitor[66]
eseroline Biomolecules 12 00694 i005Physostigmine derivativeAChE inhibitor[66]
phenserine Biomolecules 12 00694 i006Physostigmine derivativeAChE inhibitor[66]
Table 3. Chemical structures of ginsenosides [82].
Table 3. Chemical structures of ginsenosides [82].
StructureGinsenosideR1R2R3
Biomolecules 12 00694 i007PPD-typeRb1-O-Glc-Glc-H-O-Glc-Glc
Rb2-O-Glc-Glc-H-O-Glc-Ara(p)
Rc-O-Glc-Glc-H-O-Glc-Ara(f)
Rd-O-Glc-Glc-H-O-Glc
Rg3-O-Glc-Glc-H-OH
F2-O-Glc-H-O-Glc
Rh2-O-Glc-H-OH
Compound K-OH-H-O-Glc
PPD-OH-H-OH
PPT-typeRe-OH-O-Glc-Rha-O-Glc
Rf-OH-O-Glc-Glc-OH
Rg1-OH-O-Glc-O-Glc
Rg2-OH-O-Glc-Rha-OH
Rh1-OH-O-Glc-OH
F1-OH-OH-O-Glc
PPT-OH-OH-OH
Table 4. Chemical structures of ginkgolides [86,87] from GBE extracts. GBE has been described to reduce APP expression and to improve cognitive function [88,90,91,92,93].
Table 4. Chemical structures of ginkgolides [86,87] from GBE extracts. GBE has been described to reduce APP expression and to improve cognitive function [88,90,91,92,93].
NameStructureNameStructure
ginkgolide A Biomolecules 12 00694 i008ginkgolide B Biomolecules 12 00694 i009
ginkgolide C Biomolecules 12 00694 i010ginkgolide J Biomolecules 12 00694 i011
Table 5. Chemical structures and neuroprotective characteristics of plant natural products from different origin.
Table 5. Chemical structures and neuroprotective characteristics of plant natural products from different origin.
NameStructureCharacteristicsRef.
curcumin Biomolecules 12 00694 i012antioxidant, anti-inflammatory, decreases inflammation and ROS[96]
icosapent ethyl Biomolecules 12 00694 i013improves synaptic function, reduces inflammation[103]
rapamycin Biomolecules 12 00694 i014reduces Aβ deposition and pathogenic tau phosphorylation, improves synaptic plasticity, decreases neuroinflammation[107,108,109,110,111,112,113]
cannabidiol Biomolecules 12 00694 i015may protects against Aβ-induced and microglia-activated neurotoxicity in vitro, prevents hippocampal and cortical neurodegeneration, reduces tau hyperphosphorylation, regulates microglial cell migration, anti-inflammatory, antioxidant[114,115,116,117,118,119,120,121]
Table 6. Chemical structures and neuroprotective characteristics of carbohydrates from algae.
Table 6. Chemical structures and neuroprotective characteristics of carbohydrates from algae.
NameStructureSourceCharacteristicsRef.
GV971
(Sodium oligomannate)
Biomolecules 12 00694 i016marine brown algaemight act via decreasing neuroinflammation by remodeling gut microbiota and balancing the amino acid metabolism, especially phenylalanine and isoleucine[124]
porphyran Biomolecules 12 00694 i017Porphyra yezoensissuperoxide anion and hydroxyl radical scavenging activity[130]
floridoside Biomolecules 12 00694 i018Laurencia undulataanti-inflammatory activity, inhibits the production of NO and ROS, downregulates iNOS and COX-2[132]
fucoidan Biomolecules 12 00694 i019Ascophyllum nodosuminhibits ROS and TNF-α release, reduces NO, PGE2, COX-2, iNOS, MCP-1, TNF-α and IL-1β[134,135]
κ-carrageenan Biomolecules 12 00694 i020 inhibits TNF-α secretion[137]
Table 7. Chemical structures, sources and neuroprotective characteristics of lipids and peptides from algae.
Table 7. Chemical structures, sources and neuroprotective characteristics of lipids and peptides from algae.
NameStructureSourceCharacteristicsRef.
1,2-di-O-palmitoyl-3-O-(6′-deoxy-6′-sulfo-d-glycopyranosyl)-glycerol Biomolecules 12 00694 i021Scenedesmus rubescens, Scenedesmus producto-capitatus, Scenedesmus accuminatus, Scenedesmus pectinatus, Tetradesmus wisconsinensis, Eustigmatos magnusQC inhibitor[143]
1-O-palmitoyl-2-O-linolenyl-3-O-(6′-deoxy-6′-sulfo-d-glucopyranosyl)-glycerol Biomolecules 12 00694 i022QC inhibitor[143]
1-O-linolyl-2-O-palmitoyl-3-O-(6′-deoxy-6′-sulfo-d-glucopyranosyl)-glycerol Biomolecules 12 00694 i023QC inhibitor[143]
tasiamide B Biomolecules 12 00694 i024Lyngbya sp., Symploca sp.BACE-1 inhibitor[144,145]
tasiamide F Biomolecules 12 00694 i025Lyngbya sp.BACE-1 inhibitor[144]
Table 8. Chemical structures and characteristics of phenolic compounds from algae.
Table 8. Chemical structures and characteristics of phenolic compounds from algae.
NameStructureSourceCharacteristicsRef.
(−)-cartilagineol Biomolecules 12 00694 i026Laurencia dendroideaR1 = Cl; R2 = Br
AChE inhibitor
[164]
(−)-dendroidolR1 = OH; R2 = Cl
AChE inhibitor
[164]
(−)-elatol Biomolecules 12 00694 i027Laurencia dendroideaAChE inhibitor[164]
2,3,6-tribromo-4,5-dihydroxybenzyl alcohol Biomolecules 12 00694 i028Symphyocladia latiusculaAChE inhibitor, BChE inhibitor[165]
2,3,6-tribromo-4,5-dihydroxybenzyl methyl ether Biomolecules 12 00694 i029Symphyocladia latiusculaAChE inhibitor, BChE inhibitor, BACE-1 inhibitor[165]
6,6′-bieckol Biomolecules 12 00694 i030Ecklonia stoloniferadecreases of IL-6, NO, PGE2, COX-2 and iNOs[155]
Grateloupia ellipticaAChE inhibitor, BChE inhibitor, BACE-1 inhibitor[157]
8,8′-bieckol Biomolecules 12 00694 i031Ecklonia cavainhibits TNF-α, IL-1β and PGE2, downregulates iNOS and COX-2, suppresses p38 and JNK[147]
suppresses ROS, NO, PGE2, IL-6 and iNOS, inhibits NF-κB, Akt, JNK and p38 MAPK[154]
bis-(2,3,6-tribromo-4,5-dihydroxybenzyl) ether Biomolecules 12 00694 i032Symphyocladia latiusculaAChE inhibitor, BChE inhibitor, BACE-1 inhibitor[165]
dibenzol [1,4]dioxine-2,4,7,9-tetraol Biomolecules 12 00694 i033Ecklonia maximaAChE inhibitor[156]
dieckol Biomolecules 12 00694 i034Ecklonia cavainhibits TNF-α, IL-1β, PGE2 and ROS, downregulates iNOS and COX-2, suppresses p38, ERK, JNK and NO, AChE inhibitor[147,148,160]
Eisenia bicyclisinhibits NO[152]
Ecklonia stoloniferainhibits Aβ25–35 self-aggregation[163]
dioxinodehydroeckol Biomolecules 12 00694 i035Eisenia bicyclisinhibits NO[152]
Ecklonia stoloniferainhibits Aβ25–35 self-aggregation[163]
diphlorethohydroxycarmalol Biomolecules 12 00694 i036Ishige okamuraeantioxidant properties[149,150]
eckmaxol Biomolecules 12 00694 i037Ecklonia maximaprevents Aβ-induced neuronal apoptosis, decreases ROS[166]
eckol Biomolecules 12 00694 i038Eisenia bicyclisinhibits NO[152]
Ecklonia stoloniferainhibits Aβ25–35 self-aggregation[163]
Ecklonia maximaAChE inhibitor[156]
fucofuroeckol-B Biomolecules 12 00694 i039Eisenia bicyclisinhibits β-secretase, attenuates Aβ-induced cytotoxicity[167]
7-phloroeckol Biomolecules 12 00694 i040Eisenia bicyclisinhibits NO[152]
phlorofucofuroeckol A Biomolecules 12 00694 i041Eisenia bicyclisinhibits NO[152]
Ecklonia stoloniferainhibitsNO, PGE2, iNOS, COX-2, IL-1β, IL-6 and TNF-α, increases IκB-α, downregulates NFκB, JNK, p38 and Akt, inhibits Aβ25–35 self-aggregation[153,163]
phlorofucofuroeckol B Biomolecules 12 00694 i042Ecklonia stoloniferadownregulates COX-2 and NO, reduces IL-1β, IL-6 and TNF-α, inhibits NF-κB, Akt, ERK and JNK, increases IκB-α[147,151]
phloroglucinol Biomolecules 12 00694 i043Eisenia bicyclisinhibits NO[152]
Ecklonia stoloniferainhibits Aβ25–35 self-aggregation[163]
sargachromenol Biomolecules 12 00694 i044Sargassum micracanthumdecreases NO, PGE2, COX-2 and iNOS, increases IκB-α[158]
Sargassum sagamianummoderate AChE inhibitor[161]
Sargassum serratifoliummoderate AChE inhibitor, BACE-1 inhibitor[162]
sargaquinoic acid Biomolecules 12 00694 i045Sargassum siliquastrumreduces NO and iNOS, inhibits NF-κB and JNK1/2 MAPK, increases IκB-α[159]
Sargassum sagamianummoderate AChE and BChE inhibitor[161]
Sargassum serratifoliummoderate AChE inhibitor, BACE-1 inhibitor[162]
sargahydroquinic acid Biomolecules 12 00694 i046Sargassum serratifoliummoderate AChE inhibitor, BACE-1 inhibitor[162]
Table 9. Chemical structures and characteristics of isoprenoids from algae.
Table 9. Chemical structures and characteristics of isoprenoids from algae.
NameStructureSourceCharacteristicsRef.
fucosterol Biomolecules 12 00694 i047Pelvetia siliquosa, Panida australis, Hizikia fusiformis, Ecklonia stolonifera, Sargassum horridum, Undaria pinnatifidaincreases the level of antioxidant enzymes SOD, GPx and CAT, inhibits ROS production, AChE inhibitor, BChE inhibitor, BACE-1 inhibitor[152,168,169,170,171,172,173,174,175,176]
fucoxanthin Biomolecules 12 00694 i048Sargassum siliquastrum, Phaeodactylum tricornutumdecreases cytokines, prevents H2O2-induced and reduces ROS-induced DNA damage, inhibits BChE in vitro[177,178,179,180]
astaxanthin Biomolecules 12 00694 i049 decreases cytokines, inhibits nNOs, iNOS and COX-2 expression[181]
α-bisabolol Biomolecules 12 00694 i050Padina gymnosporainhibits AChE and BChE in vitro[182]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Klose, J.; Griehl, C.; Roßner, S.; Schilling, S. Natural Products from Plants and Algae for Treatment of Alzheimer’s Disease: A Review. Biomolecules 2022, 12, 694. https://doi.org/10.3390/biom12050694

AMA Style

Klose J, Griehl C, Roßner S, Schilling S. Natural Products from Plants and Algae for Treatment of Alzheimer’s Disease: A Review. Biomolecules. 2022; 12(5):694. https://doi.org/10.3390/biom12050694

Chicago/Turabian Style

Klose, Jana, Carola Griehl, Steffen Roßner, and Stephan Schilling. 2022. "Natural Products from Plants and Algae for Treatment of Alzheimer’s Disease: A Review" Biomolecules 12, no. 5: 694. https://doi.org/10.3390/biom12050694

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