Next Article in Journal
Paralytic Shellfish Toxins in the Gastropod Concholepas concholepas: Variability, Toxin Profiles and Mechanisms for Toxicity Reduction
Previous Article in Journal
The Polysaccharidic Nature of the Skeleton of Marennine as Determined by NMR Spectroscopy
Previous Article in Special Issue
Marine Sources of DHA-Rich Phospholipids with Anti-Alzheimer Effect
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Marine Natural Products against Alzheimer’s Disease: Past, Present and Future

1
Collaborative Innovation Center of Seafood Deep Processing, Key Laboratory of Aquatic Products Processing of Zhejiang Province, Institute of Seafood, Zhejiang Gongshang University, Hangzhou 310012, China
2
Hangzhou WeChampion Biotech. Inc., Hangzhou 310030, China
3
Zhejiang Chiral Medicine Chemicals Co., Ltd., Hangzhou 311227, China
*
Author to whom correspondence should be addressed.
Mar. Drugs 2023, 21(1), 43; https://doi.org/10.3390/md21010043
Submission received: 29 September 2022 / Revised: 12 December 2022 / Accepted: 30 December 2022 / Published: 5 January 2023
(This article belongs to the Special Issue Anti-Alzheimer Agents from Marine Sources)

Abstract

:
Alzheimer’s disease (AD), a neurodegenerative disease, is one of the most intractable illnesses which affects the elderly. Clinically manifested as various impairments in memory, language, cognition, visuospatial skills, executive function, etc., the symptoms gradually aggravated over time. The drugs currently used clinically can slow down the deterioration of AD and relieve symptoms but cannot completely cure them. The drugs are mainly acetylcholinesterase inhibitors (AChEI) and non-competitive N-methyl-D-aspartate receptor (NDMAR) antagonists. The pathogenesis of AD is inconclusive, but it is often associated with the expression of beta-amyloid. Abnormal deposition of amyloid and hyperphosphorylation of tau protein in the brain have been key targets for past, current, and future drug development for the disease. At present, researchers are paying more and more attention to excavate natural compounds which can be effective against Alzheimer’s disease and other neurodegenerative pathologies. Marine natural products have been demonstrated to be the most prospective candidates of these compounds, and some have presented significant neuroprotection functions. Consequently, we intend to describe the potential effect of bioactive compounds derived from marine organisms, including polysaccharides, carotenoids, polyphenols, sterols and alkaloids as drug candidates, to further discover novel and efficacious drug compounds which are effective against AD.

Graphical Abstract

1. Introduction

Dementia is an umbrella term for a particular group of symptoms. The characteristic symptoms of dementia are difficulties with memory, language, problem-solving, and other thinking skills that affect a person’s ability to perform everyday activities. Alzheimer’s disease (AD) is the most common cause of dementia, accounting for an estimated 60% to 80% of cases [1]. AD is associated with age and characterized by losing neuronal structure and function gradually [2]. Also, with the population growing and life expectancy increasing, the prevalence of AD continues to rise [3]. The symptoms of AD are mainly associated with progressive memory impairment, aphasia, purposeful complex activity ability. At the same time, Alzheimer’s sufferers have an inability to distinguish previously familiar objects through specific senses, impaired visuospatial abilities, impaired abstract thinking and numeracy, and changes in personality and behavior.
Over time, symptoms will worsen and can even be life-threatening. The pathological sequence of AD starts in the center temporal lobe firstly, which is responsible for memory, and then progresses to the regions of the frontal, temporal, parietal, motor, sensory, and subcortical areas. The neuropathology of AD is complex and not yet fully understood [4]. The 2011 National Institute on Aging and Alzheimers Association (NIA-AA) guidelines defined three phases of AD: preclinical AD, MCI (symptomatic predementia) and dementia [5]. Preclinical AD, the earliest phase from normal cognition to AD dementia is characterized by the fact that daily life and work are basically not affected and one’s ability to live independently is relatively complete (lasting about three years, accompanied by mild cognitive difficulties). The most notable deficits are memory loss (e.g., it is hard to recall newly learned knowledge, unable to communicate new messages or there are difficulties in semantic memory). Certain difficulties with coordination and planning may arise when doing some delicate motor tasks (for instance, difficulties in writing, painting or dressing also occur at this stage) [6]. Even when people walk on familiar roads, they can also lose their way [7]. In the middle stage of the disease (the MCI stage), patients have the inability to live independently and will almost die from various accidents and complications; this phase lasts about two years. Changes in behavior are more pronounced at this stage, and the patient begins to no longer recognize their family and other close people [8]. Approximately 30% of patients are at risk of urinary incontinence [9], along with muscle mass decreases, inactivity, and becoming bedridden. However, the ability to receive and send emotional signals is still there [10]. In the advanced stage of the disease, patients totally lose the ability to take care of themselves and their behaviors deteriorate; most of them pass away due to various complications within one to two years.
AD has become a serious public health issue with high medical costs and no cure. Most cases appear in people over 65 years of age, and the global morbidity is about 6%, with women having a higher morbidity rate than men. The Alzheimer’s Disease Society International’s Association published a report entitled “The Global Impact of Dementia 2013−2050”, which showed that 44 million people are ill with dementia, a 17% raise from 2009, and that the number of people worldwide with dementia is almost doubling every 20 years. By 2030 and 2050, 76 million and 135 million cases of dementia will be found around the world. China has the highest prevalence rate with about 10 million people of dementia in the world and a yearly increase of 30 percent. In China, the prevalence rate is 5% for those over 65 years old and 30% for those over 85 years old. With the fast aging of China’s population, the number of people with dementia in China will exceed 20 million by 2050, 10% of the elderly over 75 years old will have cognitive impairment, and one third of the elderly over 85 years old will have cognitive impairment [11].
In view of the increasing number of patients with AD, researchers have never stopped exploring therapeutic drug development. In recent years, many researchers have found that in the marine environment there are a large number of unique and different structures, including polysaccharides, carotenoids, polyphenols, sterol and alkaloids, which have biological and pharmacological activities [12]. Marine ecosystems cover more than 70% of the earth’s surface, accounting for about half of the global biodiversity [13]. Marine bioactive compounds have unique biological activities due to their chemical properties that are not found in terrestrial products [14]. Therefore, people are increasingly studying these resources to explore drugs that can treat human diseases [15,16]. This review briefly introduces the related pathology of AD and comprehensively expounds the drug potential of marine compounds which is found to have the potential to treat AD based on the therapeutic targets of AD. The purpose of this review is exhibiting the great potential of marine natural products, and also providing the direction for the development of new therapeutic drugs for AD.

2. Pathogenesis of Alzheimer’s Disease

According to the age of onset, Alzheimer’s Disease (AD) can be divided into two types: early onset (EOAD) and late onset (LOAD). EOAD has certain heritability, accounting for a small number of all Alzheimer’s disease. The majority of patients have LOAD. Age is the greatest risk factor for acquiring AD. In adults older than 85, the prevalence of AD is more than one in three [17,18,19]. At present, there is no final conclusion on the pathogenesis of Alzheimer’s disease. However, there is no doubt that neuroinflammation, extracellular plaques, and intracellular neurofibrillary tangles (NFTs) are key pathological trademarks of this illness. Key factors in the pathogenesis of AD are shown in Figure 1 [15]. There are several pathogenesis concerning AD as described below.

2.1. Amyloid Cascade Hypothesis

The first hypothesis is the amyloid cascade. This hypothesis concludes that Aβ aggregation as an early event in neurodegeneration occurs independently prior to tangles formation, followed by microglia and astrocyte activation, neuroinflammatory responses, oxidative stress, and other cellular molecular events [21]. The pathogenesis of amyloid protein starts from the change and cleavage of amyloid precursor protein (APP). β-secretase (BACE1) and γ-secretase change and split APP to produce insoluble Aβ raw fiber. Then, Aβ aggregation form the plaques. On the one hand, Aβ oligomerization spreads to the synaptic gap that interferes with synaptic signal transmission. On the other hand, this polymerization leads to the activation of kinases, which promotes hyperphosphorylation of microtubule correlation tau protein. Then, hyperphosphorylated protein is polymerized into insoluble NFT. After aggregation of plaque and NFT, microglia around the plaque gathered. This promotes microglial activation and local inflammatory response and aggravates neurotoxicity [22]. From the above process, it can be seen that Aβ is like a trigger button in the process of the disease, and it has been proposed that Aβ is a trigger target for pathological processes. Therefore, Aβ is a major therapeutic target [23]. Perhaps this is why many people regard it as the primary drug target.

2.2. Neuroinflammation Hypothesis

Another hypothesis is the neuroinflammation hypothesis, which is modified from the above-mentioned amyloid hypothesis. Based on this theory, the pathogenesis of AD is mainly related to a series of activities of microglia in the immune system of the central nervous system (CNS) in which that microglia continuously activate pro-inflammatory cells through the transduction of pro-inflammatory signals [24]. At this point, the microglia can’t adjust the impairment of anti-inflammatory cytokine and lipid dielectric, causing damage to the nerve degeneration and neuronal metabolites, which brings more inflammation and excessive phosphorylated tau protein increased. Long term activation of the immune response has been shown to worsen AD pathology, possibly as a result of persistent activation of microglia in a feedforward loop (termed reactive microgliosis). This leads to the accumulation of Aβ and the persistent single-shot of pro-inflammatory cytokines that start to damage neurons [25]. Many elderly people have amyloid plaques in their brains that never develop into Alzheimer’s disease. Amyloid protein accumulation itself is not enough to cause dementia. The research results of Pascoal et al. show that it is the interaction between neuroinflammation and amyloid pathology that releases the spread of tau protein, which ultimately leads to extensive brain damage and cognitive impairment [26]. Now, neuroinflammation is established as a feature of Alzheimer’s.

2.3. Ca2+ Hypothesis

The next hypothesis is the Ca2+ hypothesis. Ca2+ dysregulation is a common and pervasive pathophysiological phenomenon in AD. Zhong et al. [27] hypothesized that N-methyl-D-aspartate receptor subunit (GluN3A) is essential for sustained Ca2+ homeostasis and its deficiency is a causative factor in AD. By examining cellular, molecular and functional changes in adult/senescent GluN3A knockout (KO) mice, they concluded that chronic ‘degenerative excitotoxicity’ can lead to sporadic AD, and GluN3A is the main pathogenic factor, a lifelong moderate but sustained Ca2+ overload is a causal pathogenic mechanism of sporadic AD. Therefore, GluN3A may be an amyloid-independent therapeutic target [27].

2.4. Tau Hypothesis

The tau hypothesis is based on the notion that tau protein’s hyperphosphorylation leads to NFTs, which is one of the chief pathological conditions of AD [28]. Clinical studies have found that the quantity of hyperphosphorylated tau protein and NFTs in the cerebrum of AD patients is positively associated with the degree of clinical dementia. That is to say the higher number of hyperphosphorylated tau proteins and NFTs, the more severe clinical and stupid condition. The tau protein is encoded by the microtubule-associated protein tau (MAPT) gene on chromosome 17. MAPT produces a monolithic hydrophilic protein, which is enriched in large, naturally unexpanded regions of developing and mature neuronal axons [28]. The tau protein exists in two isomers, namely the 3-repeat sequence (3R) and 4-repeat sequence (4R), in which the 3R tau protein mainly occurs during development and 4R tau protein is mainly produced in adulthood. These two isomers maintain a balanced proportion (1:1) in the grow-up of the cerebrum. Disturbance of the proportion between 3R and 4R may lead to AD and other diseases. It is also mentioned in the amyloid hypothesis that Aβ polymerization activates kinases, which promotes hyperphosphorylation of microtubule correlation tau protein [22]. So, inhibition of kinase activity may an effective way to control tau hyperphosphorylation. Some related studies have demonstrated that reducing tau phosphorylation by inhibiting tau kinases can restore tau-dependent long-term potentiation (LTP) deficits and attenuates synaptic loss in tau transgenic mice [29]. The novel role of pathological tau protein in disease progression will provide more directions for the search of alternative disease mechanisms and related treatment strategies in the field of Alzheimer’s disease.

2.5. Cholinergic Hypothesis

The pathogenesis of AD is usually related to the decrease of neurotransmitter levels, such as serotonin, norepinephrine, dopamine, acetylcholine, etc. Acetylcholine (ACh), which is closely associated with the formation and storage of human’s memory ability, is an important neurotransmitter in the human cerebrum. Decreasing the level of Ah can directly lead to cell damage in the basal nucleus, temporal lobe, and parietal lobe, thereby reducing the level of serotonin and intensifying the development of NFT [30]. The theory states that AD-related psychiatric symptoms arise are related to the impairment of cholinergic neurons and dopaminergic transmission [31,32]. ACh regulates many key functions of the CNS by activating cell-surface receptors known as muscarinic acetylcholine receptors (M1-M5 mAChRs), which are co-expressed with D1 dopamine receptors in a specific subset of striatal projection neurons. It has been proved clinically that selective M1 agonists can improve the cognition of AD patients and reduce Aβ in cerebrospinal fluid level [33]. Besides, studies on M4 gene knockout mice showed that M4 gene deletion increased dopamine efflux in the nucleus accumbens of mice, which confirmed the physiological relevance of M4 mAChR subsets in regulating dopamine-dependent behavior and indirectly verified this hypothesis [34]. Currently, some of the drugs are used to treat Alzheimer’s disease and target acetylcholinesterase.

2.6. Glutamate Hypothesis

Finally, the glutamate hypothesis is based on the truth that cognitive disability in AD is closely related to synaptic plasticity, and that N-methyl-D-aspartate receptor (NDMAR) plays a key role [35]. Excitatory glutamatergic neurotransmission through the NMDAR is crucial for neuronal synaptic plasticity and staying alive. However, it is a potential mechanism of AD neurodegeneration that NMDAR overacts excitotoxicity and promotes cell death [36]. As a kind of AD drug, it works by inhibiting NMDAR activation based on this hypothesis exactly.

3. Currently Approved Drugs for Alzheimer’s Disease

Currently, it is reported that about 24 million people worldwide are living with AD. Alzheimer’s disease has become a public health problem, but before June 2021, there were only two types of medicines have been permitted to treat AD patients, including cholinesterase enzyme inhibitors (naturally derived, synthetic and mixed analogs) and NMDAR antagonists [37]. The pathway of acetylcholine-producing cells in AD patients is disrupted, and the role of acetylcholine inhibitory enzymes is to prevent choline enzyme from breaking down acetylcholine, thereby increasing the level of acetylcholine in the synaptic [38]. NMDAR antagonists prevent excessive activation of NMDAR, thereby preventing calcium influx and avoiding cell death and synaptic dysfunction cause by increased calcium ion concentration [36]. Although these two types of drugs have therapeutic effects on Alzheimer’s disease, they can only relieve symptoms and cannot achieve the effect of cure and prevention. In June 2021, the US Food and Drug Administration officially approved aducanumab for the treatment of Alzheimer’s disease. This is the first therapeutic drug targeting the potential pathophysiology of disease since 2003. Besides, on November 2, 2019, Sodium Oligomanne Capsules (GV-971) was approved by the State Drug Administration of China for marketing, which is the first new drug for Alzheimer’s disease that targets brain gut axis in China and the world. On April 8, 2020, Shanghai Green Valley Pharmaceuticals, China received the formal decision letter from the US Food and Drug Administration (FDA) on the Investigational New Drug (IND) application for the GV-971 international multi-center Phase III clinical study [39]. However, the manufacturer, Green Valley Pharmaceutical, announced the early termination of the international multicenter phase III clinical study of this drug on May 13, 2022, due to insufficient research funds. Table 1 presents several related information about drugs currently used for AD treatment.

3.1. Acetylcholinesterase in-jibtor (AChEI)

3.1.1. Tacrine

A common kind of drugs for curing of mild to moderate AD and related dementias is cholinesterase inhibitors (ChEIs) [40]. Tacrine was the first cholinesterase inhibitor medicine to be used, which was first synthesized in the 1930s. Tacrine has been used in patients with AD since the 1980s, having been approved by the FDA in 1993 and discontinued in 2013 [38]. Tacrine has inhibitory effects on both acetylcholinesterase (AChE) and butyrylcholinesterase (BChE), but was stopped due to its side effects such as gastrointestinal side effects and hepatotoxicity in patients.

3.1.2. Donepezil

Donepezil is a second-generation reversible AChEI drug that is also used to treat patients with mild to moderate AD [41]. Its mechanism of action is that donepezil is reversibly combined with acetylcholinesterase so that it has inhibitory effect to hydrolysis of acetylcholine. To do that, it achieved the effect of increasing the concentration of acetylcholine at the synapse in the end. However, it should be noted that donepezil is only used to treat AD symptoms, such as improving cognition and behavior, but does not change AD progression [44].

3.1.3. Rivastigmine

Rivastigmine is a reversible dual inhibitor of AChE and BChE in a slow treatment speed. There are two ways of administration: oral and transdermal. It is approved by FDA that oral dosage form of rivastigmine is used for treating mild to moderate AD. The optimum therapeutic dose is 6–12 mg/day [42]. Gastrointestinal side effects can also occur with this drug (for better tolerance it is recommended to take it with food, twice a day) [45]. Transdermal patch, another mode of administration, is a paper-thin, waterproof matrix patch, which has the advantages of easy administration and reduced tablet burden for patients taking combination drugs and patients with dysphagia. Appears to have adverse skin reactions, most commonly irritant contact dermatitis, which is usually mild, transient, and can be controlled with topical treatments (such as topical corticosteroids) [46].

3.1.4. Galantamine (GAL)

Galantamine, as a standard first-line drug, is used for treating mild to moderate AD. GAL is a selective tertiary isoquinoline alkaloid with a dual mechanism of competing with acetylcholine receptor α-subunit to prevent it combine with acetylcholine inhibitors and to activate acetylcholine [38]. Adverse reactions that may occur with this drug include weight loss, diarrhea, loss of appetite, nausea, vomiting, dizziness, headache, gastrointestinal bleeding, etc. [46].

3.2. NMDAR Antagonists

Memantine, which is an antagonist of the NMDAR, is an FDA-approved prescription drug for the treatment of moderate-to-severe Alzheimer’s disease [37]. The drug reportedly targets extrasynaptic NMDARs preferentially [47]. Adverse reactions that may occur while taking this drug include constipation, diarrhea, confusion, dizziness, and headache [46].

3.3. Aducanumab

Aducanumab, with the structural formula shown in Figure 2 [48], is one kind of monoclonal antibody of human immunoglobulin γ (Ig G)1. By recognizing Aβ epitopes, Aducanumab acts on soluble and insoluble Aβ, reduces its content, and has low affinity for Aβ monomers [48]. The most common adverse reaction of the drug in clinical trials is brain edema or hydrocephalus. The launch of aducanumab was also controversial, with only the EMERGE (NCT02484547) [49] Phase III global trial showing a significant reduction in Aβ in AD patients and a significant improvement in the CDR-SB score, a measure of cognitive function. Therefore, the FDA accelerated the approval of aducanumab but also required phase IV trials to further verify the efficacy and safety of aducanumab.

3.4. Sodium Oligomanne Capsules (GV-971)

Sodium Oligomanne Capsules (GV-971) is a low molecular acid oligosaccharide compound prepared from the extract of marine brown algae [50]. In non-clinical studies, the mechanism of action of GV-971 is to decrease peripheral inflammation that may aggravate neuroinflammation by regulating the gut microbiota, thereby reducing the neuroinflammation in the brain. GV-971 combines with Aβ in direct to decreases Aβ deposition in the brain [51,52,53]. This drug is mainly used to treat patients with mild to moderate AD and may cause skin rash, abdominal pain, and other adverse reactions in patients.

3.5. Combination Drug Therapy

Alzheimer’s disease is a chronic disease, and patients usually require long-term medication. For patients on long-term medication, therapeutic effects may be limited (in a fixed dose regimen) due to neurobiological adaptation and drug tolerance, promoting the need increase the dose further. Therefore, lowering the drug dose as much as possible while still achieving the therapeutic effect can reduce the possibility of induced toxicity. Combination drug therapy may be a good method [54]. Hafsa Amat-ur-Rasool et al. performed in vitro experiments on the inhibitory effect of anticholinesterase combined drugs on AChE protein, the inhibitory effect of AChE in differentiated neurons, and the cytotoxicity of neuronal cells. Two-drug combinations of berberine and tacrine (BerTac), berberine and galantamine (BerGal), and tacrine and donepezil (TacDon) all produce synergistic AChE inhibition results. Donepezil and galantamine (DonGal) have a synergistic effect on human AChE, but have an antagonistic effect on tcAChE. After the combination of the two drugs was applied to neuronal cells, BerTac, BerGal, DonGal and donepezil and berberine (DonBer) all showed synergistic inhibition of AChE. BerGal produced the most potent synergy, reducing the total drug dose by 72% [55]. In China, there are also a few reports on the use of sodium oligomannate or memantine combined with donepezil in the treatment of mild Alzheimer’s disease cases.
Currently, there is no specific drug that can treat AD. Therefore, the development of new specific drugs is an urgent problem to be solved. The huge number of marine organisms can produce a large number of natural products with unique structures, some of which have neuroprotective effects and may be used for AD treatment. At the same time, the successful marketing of GV-197, which was made from marine brown algae extracts in China, shows us the bright prospect of marine biological ingredients as the treating drugs of AD.

4. Anti-AD Marine Natural Products

Marine natural products have increasingly allured great scientific interest since they can show significant biological activity 10 times higher than terrestrial organisms [14]. Through the study of the biological functions of marine natural products, it has continuously proved its potential in the biomedical neighborhood, including anti-thrombotic, anti-coagulant, anti-inflammatory, anti-hypertension, anti- diabetes, heart protection, and neuroprotection effects [56]. A large and growing literature has investigated that marine natural products have great potential against AD due to their chemical compounds, including polysaccharide, polyphenols, sterol, carotenoids, diterpenoids, alkaloids, amino sulfonate, marine toxins and macrolide-type compounds [14], as summarized in Table 2. In the process of extracting marine compounds and developing new drug molecules, several key parameters help to determine the overall safety of precursor drugs, including absorption, distribution, metabolism, excretion, and toxicology (ADMET). We have listed the information related to trans-blood brain screen, skin permeability, and drug toxicity of some marine compounds in Table 3.

4.1. Polysaccharides

Polysaccharides are polymeric carbohydrates formed by the combination of sugar chains and glycosidic bonds with at least ten or more monosaccharides with high molecular weight [73]. So far, more than 300 natural polysaccharide compounds have been founded in the literature, such as fucoidan, chitosan, mannan, and seaweed polysaccharide, among which the fucoidan and chitosan have the potential to treat AD.
Fucoidan (Figure 3A) is a sort of complicated polysaccharide compound, which separated from brown seaweed and some marine invertebrates [74]. It mainly contains L-fucoses and sulfate groups [75], which has extensive pharmacological actions, such as anti-virus, oxidation resistance, anti-inflammatory, and anticoagulant, improving Aβ1-40 induced memory impairment, etc. [75,76]. Fucoidan consists of two chain structures, one with α- (1→3) connected L-fucose residues as the chain and the other with alternating α- (1→3) and α- (1→4) connected L-fucoses that may be sulfate substituted on C-4 [3]. It can block the activation of the enzymes caspase-9 and caspase-3 by β-amyloid protein (Aβ), which plays an important role in apoptosis processes. It is precisely due to the fact that it can inhibit apoptosis that the compound also has a significant effect in terms of neuronal death suppression. A fucoidan treatment has been found capable of suppressing the intracellular Ca2+ responses of neurons [77]. The superfluous accumulation of intracellular Ca2+ may go in front of the wound and put the neurons to death. Studies have indicated that high attention of fucoidan can cause and prevent the dying of dopaminergic neurons in vivo and in vitro models [57,78]. Fucoidan can reduce the neurotoxicity of β-amyloid protein, since it can reduce toxicity by reducing the inhibition of β-amyloid protein on protein kinase C (PKC) phosphorylation [57,79]. So, it can produce neuroprotective effect. In addition, it was found that fucoidan could reduce neuronal damage in AD mouse model [80]. The compound can raise mitochondrial activity and reduce the release of lactate dehydrogenase (LDH) and reactive oxygen species (ROS) for enhancing the neuroprotection. It can enhance the expression of anti-apoptotic-protein Bcl-2 and lessen the expression of Pro apoptotic protein Bax for dramatically inhibiting neuronal apoptosis [81].
Chitosan (Figure 3B), called deacylated chitin, is a natural polysaccharide that can be obtained from marine crustaceans or in the cell walls of some fungi. Chitosan is a polymer of β- (1→4)-D-glucosamine or a copolymer with N-acetyl-D-glucosamine, which is formed by enzymatic N-deacetylation of chitin on fungal cell wall by chitin deacetylase [82]. Chitosan is difficult to be used in food and biomedical fields due to its poor solubility [83]. Therefore, researchers mostly convert chitosan to chitosan oligosaccharides (COS) that are easily soluble in water [84]. Chitosan and COS have been proven to plays a significant part in oxidation resistance and antibacterial, anti-apoptosis, and immune regulation [58,85]. Khodagholi et al. [86] studied the effect of chitosan on NT2 neuron cells induced by H2O2 and FeSO4, and found that chitosan treatment could inhibit the death of NT2 neuron cells induced by oxidative stress. Moreover, the formation of Aβ in NT2 neurons pretreated with chitosan was significantly lower than that in control cells. The results showed that the Aβ level could controlled by treatment with this chitosan, suggesting that chitosan had a certain protective effect on AD. Yoon et al. synthesized COS derivatives with different substituents. Among the three COS derivatives, Diethylamino-ethyl COS (DEAE-COS) has the strongest AChEIs activity [87]. The results showed that DEAE-COS was identified as a competitive AChEIs and that chitosan and its derivatives could inhibit the activity of acetylcholinesterase and had a potential neuroprotective effect on neuronal disorders [58]. Numerous studies have shown that it can restrain the formation of amyloid-β and the activities of acetylcholinesterase. Also, it has the peculiarity of anti-neuroinflammation and anti-apoptosis properties [58,88]. So far, the neuroprotective effects of chitosan and its derivatives have mostly been observed in vitro [89]. Therefore, further studies are needed to investigate their activity in mouse model systems and/or human subjects. Taken together, these results revealed the potential of chitosan and its derivatives as potential therapeutic candidates for neurodegenerative diseases, and they hold great promise in future drug development.
Mannan (Figure 3C) is a polysaccharide widely existing in nature, which exists in plants, marine bacteria, yeast, etc. It has the characteristics of high water solubility, stability, and viscosity [90]. Mannan is composed of monosaccharide D-mannose linked by β-1,4-mannoside linkage [91]. Liu et al. [59] investigated the effect of the hydrolytic product of mannan, mannan oligosaccharides (MOS) on 5xFAD-Tg mice, a classic transgenic AD mouse model. It was found that after eight weeks of MOS (0.12%, w/v in the drinking water) oral administration, cognitive function and spatial memory were significantly improved and behavioral disorders were reduced in mice. In addition, immunofluorescence staining showed that MOS treatment improved the neuronal morphology of AD-affected mice brains, and greatly reduced the accumulation of Aβ in the cerebral cortex and hippocampus. Importantly, MOS treatment reduced the overexpression of tumor necrosis factor-α (TNF-α) and interleukin 6 (IL-6) in the brains of AD mice, effectively alleviated neuroinflammation and oxidative damage in the brains of AD mice [59]. Therefore, further clinical trials are needed for evaluation to explore the therapeutic effect of chitosan on AD patients.

4.2. Polyphenols

Polyphenols are a sort of neuroprotective compound that consist of aromatic ring and one or more phenolic rings in natural plants [92,93]. Algae are rich in polyphenols (especially phloroglucinol, a kind of phlorotannins which has been proved to have many biological activities) [94]. An experiment was conducted to investigate the effect of brown algae polyphenols on scopolamine-induced memory impairment in mice. The results showed that phlorotannins (50 or 100 mg/kg) orally supplemented for six weeks improved scopolamine-induced memory impairment in behavioral tests [93]. Phlorotannins decreased acetylcholinesterase activity in the brain, and significantly reduce lipid peroxidation levels but increased glutathione levels and superoxide dismutase activity [93]. In addition, phlorotannins upregulate the expression levels of brain-derived neurotrophic factor (BDNF), phosphorylated extracellular signal-regulated kinase (ERK), and cyclic AMP-response element-binding protein (CREB), which influence cholinergic dysfunction and memory deficit [95,96]. It has been found that phlorotannins including dieckol, eckol, and 8,8′-bieckol can inhibit the BACE-1 enzyme.
Currently, drugs used to treat AD work mainly by inhibiting acetylcholinesterase, so the discovery that phlorotannins including dieckol, eckol, and 8,8′-bieckol have an inhibitory effect on BACE-1 enzyme may improve medication and treatment for AD patients [60,97]. Liu et al. evaluated the neuroprotective effects of three polyphenols, including 8,8’-bieckol, dieckol, and eckol, on Aβ25-35-mediated cytotoxicity in PC12 cells. Of these, dieckol (Figure 3D) showed the greatest protective effect, although both had been shown to inhibit inflammatory responses by inactivating the NF-κB pathway [98]. Dieckol has been found to have a powerful inhibitory effect on amyloid-β peptide (Aβ) accumulation in and out of the cell. Yoon et al. [60] demonstrated for the first time that dieckol regulates APP proteolysis and Aβ production by regulating the phosphoinositide 3-kinase (PI3K)/protein kinase B (Akt)/Glycogen synthase kinase 3β (GSK-3β) signaling pathway. In addition, the addition of PI3K inhibitor LY294002 counteracted all of the effects of dieckol, suggesting that Akt/GSK-3β is the main pathway mediating Aβ production by Swedish mutant APP over-expressed Neuro-2a (SweAPP N2a) cells. The current findings support a better understanding of the important role of dieckol in the prevention of AD and its potential as a promising source of anti-AD drugs [60,99].

4.3. Sterols

Marine sterols are sterol compounds produced by organisms in ocean. They are the commonest natural organic compounds in the ocean (and are mostly from marine algae). Marine sterols such as fucosterol (Figure 3E) have the capability to lessen the content of cholesterol in the blood, leading to the tubular formation of cholesterol in the body [100]. However, the out of balance of cholesterol homeostasis will lead to inflammation, which is related to the pathobiology of neurological diseases [61]. Furthermore, fucosterol exhibited moderate anti-AChE activity by Ellman’s method [101]. So, it may have an improvement effect on the symptoms of AD. In addition, sterols have anti-inflammatory effects in macrophages and can prevent LPS- or Aβ-mediated neuroinflammation [102]. It is found that sterols have anti-BACE1, and Aβ aggregation inhibitory activities [62]. Castro-Silva et al. evaluated the AchE inhibitory activity of sterols isolated from Sargassum seagrass extracts in vitro and in silicon. The sterol not only has a higher affinity for AchE compared to the positive control, but is also a non-competitive human acetylcholinesterase (hAChE) inhibitor, which differs from the stabilizing effect of galantamine (competitive hAChE inhibitor) [103]. Marine sterols can reduce cell apoptosis and inhibit neuroinflammation by regulating brain-derived neurotrophic factor (BDNF, synaptic growth factors associated with memory and learning), nuclear factor erythroid 2-2-related factor 2 (Nrf2), and nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) signaling systems [98,104].

4.4. Carotenoids

Carotenoids are a kind of significant pigments in natural world, which extensively exist in animals, plants, fungus, and algae (the vast majority of which come from algae) [64,105]. They are formed by linking isoprenoids units and exist in various colors, including fucoxanthin, lycopene, lutein, astaxanthin (Figure 3G), etc. In fact, fucoxanthin (Figure 3F), which was abundant in brown seaweed, was isolated from S. horneri and evaluated for its effect on cognitive impairment in vivo and its ability to inhibit some key enzymes in vitro by Lin et al. [106]. It was concluded that fucoxanthin could effectively reverse scopolamine-induced cognitive impairment in mice, significantly increased choline acetyltransferase (ChAT) activity and BDNF expression, and decreased AChE activity in scopolamine treated mice, suggesting that fucoxanthine has the potential to enhance cognition [12,106]. Another study explored the direct effect of fucoxanthine on Aβ assembly and evaluated the effect of fucoxanthine on the cognitive performance of Aβ oligomer-treated mice [107,108]. The results demonstrated that fucosanthin may effectively reverse cognitive impairment by inhibiting oxidative stress and upregulating the expression of BDNF and ChAT in Aβ1-42 oligomeric-treated mice, suggesting that fucosanthin reduces Aβ neurotoxicity in vivo, which may be useful for prevention of AD [107]. In addition, fucoxanthin also showed anti-inflammatory and antioxidant effects [63]. Lycopene (LYC,Figure 3G), a fat-soluble carotenoid, was found to have inhibitory effects on LPS-induced memory loss, to inhibit the phosphorylation of lipopolysaccharide (LPS)-treated BV2 microglia MAPKs and NF-κB, and activate the Nrf2 signaling pathway in the Morris water maze test of mice, suggesting that LYC may be a preventive strategy for neuroinflammatory related diseases such as AD [65]. In vitro and in vivo studies were conducted in rat models to investigate the effects of astaxanthin on lipopolysaccharide LPS-induced inflammatory responses [109]. The results showed that the anti-inflammatory effect of astaxanthin (100 mg/kg) was higher than that of commonly used anti-inflammatory drugs (10 mg/kg), and that LPS-fed mice treated with astaxanthin showed a dose-dependent anti-inflammatory effect. Astaxanthin plays an anti-inflammatory role by inhibiting the production of (nitric oxide) NO, prostaglandin E2 (PGE2), TNF-α and interleukin-1β (IL-1β) [64]. There are currently clinical trials (NCT05015374) of astaxanthin in patients with AD taking astaxanthin or a placebo. The trials are expected to end in 2024, after which the possible benefits of astaxanthin on Alzheimer’s disease will be examined by studying patients’ mental status, cognitive ability, and clinical dementia scores. Overall, the results of the neuroprotective effects of carotenoids in vitro and in vivo are encouraging, but further clinical studies in humans are needed to draw conclusions about the full potential for treating neurodegenerative diseases such as AD.

4.5. Diterpenoids

Diterpenes are terpenoids with four isoprene units which broadly exist in terrestrial and marine organisms [110]. Researchers have found that the compounds of the gracilins, structural analogue gracilin A (Figure 3H) in sponges can inhibit BACE-1, a kinase regulated by extracellular signals, and reduce hyperphosphorylation of tau protein [111], which has a neuroprotective effect on primary neurons [112]. In addition, they can also express strongly Nrf 2 involved in the activation of antioxidant pathway, produce important anti-inflammatory effects, and reduce the production of ROS induced by amyloid-β [88]. It can also induce targeting mitochondria to inhibit mitochondrial oxidation and play a neuroprotective role [113]. A study showed that Gracilin A could protect SH-SY5Y cells from hydrogen peroxidation-induced damage by reducing reactive oxygen species (ROS) levels, restoring glutathione (GSH) content, improving mitochondrial membrane potential (MMP) and increasing cell survival rate [114]. In a different study, the activity of Gracilin A derivatives was evaluated, including their ability to regulate antioxidant gene expression in SH-SY5Y cells and their anti-neuroinflammatory potential in LPS-activated BV2 microglia [115]. The results showed that Gracilin A can regulate the translocation of Nrf2 and NF-κB and reduce the activation of p38 mitogen-activated protein kinase (p38 MAPK) in SH-SY5Y and BV2 cells [115,116].In addition, it was also found that the abietane diterpenoids also have neuroprotective effects from Phlegmariurus carinatus [117]. Harziane Diterpenes from Deep-Sea Sediment Fungus Trichoderma sp. has a potential anti-inflammatory effect [117].

4.6. Alkaloids

Alkaloids are a kind of alkaline compounds containing nitrogen that exist in nature, mainly in plants, animals and algae. In the ocean, marine macrocyclic alkaloids mainly manzamines (Figure 3I), 3-alkylpiperidines, 3-alkyl pyridinium salts and so on [118]. Manzamine alkaloids are a class of complex β -carbonyl alkaloids isolated from spongy bodies with unique nitrogenous polycyclic systems [119]. It was found that the manzamine alkaloid is an inhibitor of glycogen synthetase kinase-3 beta (GSK3β) and reduces the activity of GSK3β through the hydrogenation of C-32/C-33 double bond and the oxidation of C-31 to ketone [67,120]. In cell experiments, manzamine A showed strong inhibition of tau phosphorylation in cells [67]. When the manzamines were evaluated in trials related to nervous system function and pathology, it did not show any effect on AChE or β-secretase, nor did it show significant ability to protect human neurocytoma SH-SY5Y cells against oxidative stress-induced cell death [121]. It is thus known that this alkaloid can be used to restrain the formation of NFTs by the hyperphosphorylation of tau protein, but has no effect on other clinical features of AD [122] Indole alkaloids (Figure 3J) are a kind of macromolecular compounds isolated from Streptomyces sp. In vivo, it has high activity for transcription of Nrf2 and has a neuroprotective effect [68]. Indoles have been confirmed to reduce tert-butyl hydroperoxide (TBHP)-induced cell death, demonstrating their neuroprotective potential and having little cytotoxic effect on human neuroblastoma SH-SY5Y and microglial BV2 cells [123]. It can be seen that this alkaloid can be considered to inhibit the occurrence of intracellular NFTs caused by tau hyperphosphorylation, but has no effect on other clinical features of AD, and is a potential drug for the treatment of AD. Alkaloids have been proven to inhibit Aβ plaque production and the hyperphosphorylation of tau protein, inhibit neuroinflammation and reduce apoptosis, activate autophagy, and reduce potential risk factors for AD, so they can have the potential to become a lead in AD treatment [124].

4.7. Amino Sulfonates

Amino sulfonate compound is a kind of compound matter that takes the place of hydroxyl group in sulfuric acid and belongs to the sulfur containing protein. For example, homotaurine (Figure 3K), obtained from various marine red algae, which is a kind of natural amino sulfonate compound. It has been confirmed that it has a protective effect on neurons in vitro and in vivo models in clinical studies [125]. The homotaurine reduces the toxic effects on neurons by encapsulating amyloid peptide to reduce misfolding and aggregation of amyloid to prevent the formation of amyloid-β oligomers and neurotoxic compounds [69]. The amyloid -β oligomer is a key pathogenic factor in neurological diseases and can form Aβ-fibrils and protofibrils that ultimately lead to the formation of amyloid plaques [70]. On the other hand, homotaurine has a similar structure to the neurotransmitter γ-aminobutyric acid (GABA), so it can specifically act on GABA receptors to reduce neurotoxic damage induced by glutamate delivery [126,127,128]. In a Phase II clinical trial, homotaurine has been shown to safely reduce the concentration of Aβ42 in cerebrospinal fluid (CSF) in patients with mild to moderate AD, contributing to the potential for improvement in the disease [116]. However, homotaurine was evaluated in phase III trials (ALPHASE trial) of mild-moderate AD and showed insufficient clinical efficacy [69]. The results showed a positive trend of Clinical Dementia Rating-Sum-of-boxes (CDR-SB), and the Alzheimer’s Disease Assessment Scale-cognitive subscale (ADAS-cog) results showed a statistically significant difference, but the planned analysis of the psychometric data showed no statistically significant difference [69]. Nevertheless, only the secondary endpoints of the study showed neuroprotective effect inhibiting Aβ activity of homotaurine. Therefore, further evidence is needed to determine the effectiveness of homotaurine in the improvement of AD, providing new possibilities for the treatment of AD.

4.8. Marine Toxins

Marine toxins are naturally toxic chemicals from the ocean environment [129]. Nereistoxin (NFX) was found to have a strong binding affinity with nicotinic acetylcholine receptor (nAChR) in mouse brain [130], as was Spirolides. Spirolides (SPX) are cycloimine lipophilic marine toxins, mainly produced by Alexandrium ostenfeldii (A. ostenfeldii) [71]. The toxicity of these compounds was detected for the first time in bivalve mollusks when mice that were intraperitoneally injected with scallop and mussel lipophilic extracts died unusually quickly, indicating a strong toxic reaction in the mice [131]. The mechanism of action of SPXs was not fully understood, but it had been proposed that cholinergic receptors are the main sites of action of these toxins [132]. It was non-toxic to humans and is an antagonist of nicotinic choline receptors, enhanceing the expression of the choline acetyltransferase enzyme (ChAT) [133]. The main representative substance was 13-desmethyl spirolide C (13-desMeC) (Figure 3L) which had been shown to reduce β-amyloid accumulation and reduce tau hyperphosphorylation through its action on cholinergic receptors in a three-transgenic mice (3xTg) neuron model [134]. This was due to the fact that the treatment of 3xTg cortical neurons with the toxin can observably decrease the levels of the hyperphosphorylated isoforms of tau recognized by AT8 and AT100 antibodies and the levels of intracellular β-amyloid [135]. It could also eliminate the glutamate-mediated neurotoxic effects, decreasing intracellular accumulation of Aβ and phosphorylated tau levels in neurons in vitro. Thus, it could be used for new treatment of AD in the body from the view of barrier of blood-brain [55]. Anatoxin a(s), a marine toxin originally extracted from filamentous cyanobacterium Anabaena flos-aquae, had been extensively studied for its AChE inhibition potential [136]. In vitro, the activity of AChE was determined by colorimetric assay. The results showed that anatoxin a(s) had a non-competitive inhibitory effect on AChE. These results were confirmed in vivo, where rats treated with anatoxin a(s) showed similar signs of anticholinesterase poisoning [137]. Therefore, marine toxins have the potential to treat AD by targeting certain targets, which requires further animal trails and clinical trials.

4.9. Macrolide-Type Compound

The macrolide caniferolide A (Figure 3M) is a kind of macrolide-type compound extracted from Streptomyces caniferus, a actinomycete in the ocean, with potential biological activity [138]. Caniferolide A have been confirmed to have neuroprotective effects [139]. Caniferolide A can reduce the content of neuroinflammatory markers in lipopolysaccharide-activated BV2 glial cells, block the transfer of FFKB-P65 to the nucleus and activate the Nrf2 pathway [138]. Rebeca Alvariño et al. [139] examined tau phosphorylation in SH-SY5Y tau441 cells and found that caniferolide A can reduce Thr212 and Ser214 phosphorylation by targeting p38 and c-Jun N-terminal kinase (JNK) mitogen activated kinases (MAPK). In addition, the antioxidant activity of macrolides was determined in the oxidative stress model of SH-SY5Y cells treated with H2O2. It was found that the compound could decrease the release levels of proinflammatory cytokines (IL-1β, IL-6 and TNF-α), ROS and NO, and increase the cell viability and GSH content. Finally, in order to confirm the anti-inflammatory effect of caniferolide A, BV2 microglia and SH-SY5Y neuroblastoma cell lines were trans-well co-cultured. The results showed that caniferolide A significantly increased the survival rate of neuroblastoma cells, confirming its neuroprotective properties [139]. All of these indicated that the compound could be effective and possibly used for novel AD treatment for AD.
Bryostatin 1 is a macrolide compound with high oxygen structure, isolated from the marine invertebrate Bugula neritina, which can be used as an effective regulator of protein kinase C (PKC) [72]. It can bind to PKC, cause quick short activation and autophosphorylation of PKCs, and induce continuous translocation of PKC membrane and continuous downregulation of PKC, leading to increased production and release of BDNF in the central nervous system [140]. Intraperitoneal injection of bryostatin 1 activates PKCε in the brain and prevents Aβ elevation, synaptic loss, and memory deficits in AD mouse model [141]. Preliminary safety and tolerability data for Bryostatin-1 in AD have been evaluated in phase IIa clinical trials (ClinicalTrials.gov identifier NCT02221947) with no serious adverse events and positive results [142]. Bryostatin concentration in the blood peaked one hour after the patient was infused. However, long-term treatment with bryostatin can induce downregulation of PKCε, depending on the duration and dose level of treatment. In another Phase II clinical study, bryostatin showed better efficacy, tolerability and safety when used to improve cognitive loss in 150 patients with advanced AD. However, the initial improvement was not significant in the full analysis set (FAS), and in the completer analysis set (CAS), primary and secondary analyses showed positive results in the bryostatin (20 μg) treatment group compared to the placebo group [143]. Therefore, these analyses collectively suggest that further trials are needed to evaluate the role of bryostatin in cognitive function in patients with AD. Therefore, in past few years, these compounds have been identified as promising drugs with potential.

5. Challenges and Opportunities in Developing Marine Natural Products for Alzheimer’s Disease

In 2018, the Alzheimer’s Disease International estimated that around 50 million people worldwide have dementia, a figure expected to triple by 2050, with two-thirds living in low-and middle-income countries [144]. Alzheimer’s disease has become a huge challenge the world is facing. However, unlike the rapidly growing number of Alzheimer’s patients, the types of drugs used to treat Alzheimer’s have not increased significantly. One of the reasons for the slow renewal of therapeutic drugs is the long clinical cycle of drugs. For instance, the phase 3 trial of Aducanumab (NCT05310071) lasted three years, and the phase Ⅲ trial of GV-971 lasted five years (NCT04520412). The second reason is the high failure rate of clinical trials. However, AD drugs have a high clinical failure rate, including several pharmaceutical giants (such as Pfizer, Roche, etc) [145]. Most clinical drug trials target at Aβ and tau, but the clinical failures of the recent decades indicate that there are further pathological mechanisms [146]. This prompted researchers to focus on the development of multi-targeted drugs [122,146,147]. Multi-target targeted ligands (MTDLs) strategy has been proposed and developed many times. The MTDLs design strategy involves the incorporation of distinct pharmacophores of different drugs in the same structure to get hybrid molecules [148]. The most widely used method is combining the structure of the FDA approved cholinesterase inhibitor with another drug with biological characteristics that is useful for treatment or combined using of several drugs. A new tacrine derivative with acetylcholinesterase inhibition (AChE) and brain-derived neurotrophic factor (BDNF) activation was obtained by linking tacrine with a fragment of huperzine A, and it has been shown to have a cognitive enhancement effect in two kinds of AD mice (APP/PS1transgenic mice and β-amyloid (Aβ) oligomers-treated mice) without inducing significant hepatotoxicity [149]. However, no relevant clinical trial information was found. A clinical trial (NCT01362686) involving a combination of three commonly drugs (donepezil, galantamine, rivastigmine) for AD was terminated due to a low study accrual.
The development of MTDLs also provides opportunities for the development of drugs with marine compounds as the main components. It can be seen from Table 2 that the targets for AD of most marine compounds are multidirectional rather than single. A single compound can act on multiple targets, which means that compared with synthetic drugs, the cumbersome drug synthesis process is eliminated. Secondly, marine compounds themselves are derived from edible natural organisms, many of which are edible compounds and have been widely used in food, cosmetics and other industries, indicating that their use safety is guaranteed [91,150,151]. At present, the biggest limitation of marine compounds to become clinical drugs is the lack of clinical trials. Most of the reported marine compounds with potential to treat Alzheimer’s disease remain in the animal experiment stage. As for the clinical progress of marine compounds mentioned in Section 4, only bryostatin (NCT02221947) and astaxanthin (NCT05015374) were found to have related clinical trials. Many marine compounds have not been put into clinical trials. This situation may be related to a lack of collaboration between different disciplines, especially within medical science. Analyzing the references, it was found that most of the authors who proposed marine compounds with potential to treat AD were involved in the chemical, biological, or marine sciences, and there was generally a lack of authors with medical backgrounds. The lack of support from medical professionals has greatly slowed the progress of marine compounds to the clinical platform. At present, the biggest limitation of marine compounds to become clinical drugs is the lack of clinical trials. Most of the reported marine compounds with potential to treat Alzheimer’s disease remain in the animal experiment stage. The Center for Disease Control and Prevention (CDC) could take the lead in drafting a protocol and recruiting medical researchers as well as chemical, biological, and marine researchers, which would accelerate clinical research of marine drugs. Another challenge for marine drugs is the stability and extraction of marine compounds. For example, marine indole alkaloids have many biological activities and their neuroprotective properties can be used in the treatment of AD. However, the determination of bioactivity has become a difficult problem due to the insufficient amount of isolated and purified compounds [68]. In another case, the yield of phlorotannins extracted from Sargassum fusiforme can only reach 6.36% [152]. Translating chemical diversity into pharmacological diversity is big problem, so the isolation and extraction of compounds in the drug application development process is big challenge [68]. It is gratifying that there is a lot of research focused on the green extraction of natural compounds [137,153,154,155]. In addition, food level nanocapsules, encapsulation technology and drug target delivery technology have gradually matured [156,157,158,159], which provide a strong technical guarantee for achieving transformation from marine compounds to marine drugs. So, there is no need to worry too much about this issue. With in-depth study, these problems can be solved. Furthermore, over the past decade, several efforts have been made to discover new biomarkers that could enable more accurate and rapid diagnosis of neurodegenerative diseases. These biomarkers include magnetic resonance imaging (MRI), which targets the cerebral cortex, white matter, etc., positron emission tomography (PET), which analyzes tau lesions and beta-amyloid accumulation, and cerebrospinal fluid (CSF) targeting β-amyloid polypeptides, β-amyloid oligomers, and tau peptides as well as blood biomarkers [160,161]. Biomarkers contribute to more rapid and accurate diagnosis, provide an indication of disease progression, and identify the best drug for a particular individual. Integrating these biomarkers into drug development or clinical trials for neurodegenerative diseases is an important step to help develop and demonstrate drug efficacy and target involvement [162]. The discovery of more drug targets will certainly encourage more research teams to explore new drugs, which also provides opportunities for marine compounds.
Therefore, it can be seen that marine resources offer us a huge library of potential research drugs. Marine natural molecules can be used as lead compounds for the development of drug candidates against AD. In order to ensure that drug development has a promising research progress and provide good opportunities and application prospects in the biomedical field, it is necessary to further exploration of molecular mechanisms, toxicity and side reaction of active ingredients, along with the development of in vivo and in vitro researches, are vital to the development of novel drugs for the treatment of AD, and we need describe their health implications [163]. In conclusion, using marine compounds for treating neurodegenerative diseases is both an opportunity and a challenge.

6. Remarks and Future Perspectives

This review examines the pathogenesis of Alzheimer’s disease in brief and emphatically introduced various marine compounds with the potential to treat Alzheimer’s disease. The literature shows that marine compounds are a cost-effective and environmentally friendly resource with high biomedical potential. Sodium oligomanne capsules (GV-971), an acidic oligosaccharide compound prepared from the extract of marine brown algae, was successfully applied in clinical practice, which also indicates the drug value of marine natural compounds. Nevertheless, the development of potential drugs is mostly undertaken by academics or small biotechnology laboratories. So, it requires the joint efforts and cooperation of researchers from all disciplines, the active advocacy of government departments, and the trust and cooperation between doctors and patients to accelerate the progress of marine compounds from laboratory to clinical trials. In addition, in order to search for anti-AD compounds of marine origin with significant neuroprotective activity and better understand the potential advantages of the characteristics of various marine organisms and their compounds for the healthy of people, it is necessary to encourage and stimulate more investment in biotechnology to contribute to the more sustainable development and use of these marine resources in the future.

Author Contributions

Investigation, methodology, writing original draft preparation, D.H., Y.J. and X.H.; data curation, investigation, methodology, Y.Z., D.C., J.T., Q.C., C.S., J.Y. and M.W.; conceptualization, writing—review and editing, Y.Z., H.Z. and Y.L.; supervision, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Hangzhou Biomedical and Health Industry Development Support Project (2022WJC058 and 2022WJC122) and Department of Science and Technology of Zhejiang Province (No.LGN18C200016).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 2021 Alzheimer’s Disease Facts and Figures—2021—Alzheimer’s & Dementia—Wiley Online Library. Available online: https://alz-journals.onlinelibrary.wiley.com/doi/10.1002/alz.12328 (accessed on 7 December 2022).
  2. Yan, X.; Hu, Y.; Wang, B.; Wang, S.; Zhang, X. Metabolic Dysregulation Contributes to the Progression of Alzheimer’s Disease. Front. Neurosci. 2020, 14, 530219. [Google Scholar] [CrossRef]
  3. Ballard, C.; Gauthier, S.; Corbett, A.; Brayne, C.; Aarsland, D.; Jones, E. Alzheimer’s Disease. Lancet 2011, 377, 1019–1031. [Google Scholar] [CrossRef] [PubMed]
  4. Cabrera, M.J.A.; Pérez, R.M.; Ravelo, A.G.; Rodríguez, D.H.; Davison, G.P. Patogenia y tratamientos actuales de la enfermedad de Alzheimer. Rev. Cuba. Farm. 2014, 48, 508–518. Available online: https://www.scienceopen.com/document?vid=85802991-f11a-4e69-8f79-6cc467306d4d. (accessed on 20 August 2022).
  5. Albert, M.S.; DeKosky, S.T.; Dickson, D.; Dubois, B.; Feldman, H.H.; Fox, N.C.; Gamst, A.; Holtzman, D.M.; Jagust, W.J.; Petersen, R.C.; et al. The diagnosis of mild cognitive impairment due to Alzheimer’s disease: Recommendations from the National Institute on Aging-Alzheimer’s Association workgroups on diagnostic guidelines for Alzheimer’s disease. Alzheimers Dement. 2011, 7, 270–279. [Google Scholar] [CrossRef] [Green Version]
  6. Neuropsychological Assessment, 5th Edition—ScienceOpen. Available online: https://www.scienceopen.com/document?vid=37ea5eff-64eb-42b1-95b4-b5de983e1424 (accessed on 26 September 2022).
  7. Alzheimer’s Disease Clinical and Research Update for Health Care Practitioners. Available online: https://www.hindawi.com/journals/jar/2013/207178/ (accessed on 26 September 2022).
  8. McKhann, G.M.; Knopman, D.S.; Chertkow, H.; Hyman, B.T.; Jack, C.R.J.; Kawas, C.H.; Klunk, W.E.; Koroshetz, W.J.; Manly, J.J.; Mayeux, R.; et al. The diagnosis of dementia due to Alzheimer’s disease: Recommendations from the National Institute on Aging-Alzheimer’s association workgroups on diagnostic guidelines for Alzheimer’s disease. Alzheimers Dement. 2011, 7, 263–269. [Google Scholar] [CrossRef] [Green Version]
  9. The Neuropsychological Profile of Alzheimer Disease. Available online: https://perspectivesinmedicine.cshlp.org/content/2/4/a006171 (accessed on 26 September 2022).
  10. Achterberg, W.P.; Pieper, M.J.; van Dalen-Kok, A.H.; De Waal, M.W.; Husebo, B.S.; Lautenbacher, S.; Kunz, M.; Scherder, E.J.; Corbett, A. Pain Management in Patients with Dementia. Clin. Interv. Aging 2013, 8, 1471–1482. [Google Scholar] [CrossRef] [Green Version]
  11. Jin, X.P.; Shen, M.; Vasculocardiology, D.O. Research Progress of Continuing Nursing Care for Patients with Hypertension after Discharging from Hospital. Chin. J. New Clin. Med. 2017, 9, 933–936. Available online: http://en.cnki.com.cn/Article_en/CJFDTotal-ZYLN201709033.htm (accessed on 26 September 2022).
  12. Ghoran, S.H.; Kijjoa, A. Marine-Derived Compounds with Anti-Alzheimer’s Disease Activities. Mar. Drugs 2021, 19, 410. [Google Scholar] [CrossRef]
  13. Hosseini, S.F.; Rezaei, M.; McClements, D.J. Bioactive Functional Ingredients from Aquatic Origin: A Review of Recent Progress in Marine-Derived Nutraceuticals. Crit. Rev. Food Sci. Nutr. 2020, 62, 1242–1269. [Google Scholar] [CrossRef]
  14. Bălașa, A.F.; Chircov, C.; Grumezescu, A.M. Marine Biocompounds for Neuroprotection—A Review. Mar. Drugs 2020, 18, 290. [Google Scholar] [CrossRef]
  15. Altmann, K.-H. Drugs from the Oceans: Marine Natural Products as Leads for Drug Discovery. Chim. Int. J. Chem. 2017, 71, 646–652. [Google Scholar] [CrossRef] [PubMed]
  16. Khotimchenko, Y.S.; Silachev, D.N.; Katanaev, V.L. Marine Natural Products from the Russian Pacific as Sources of Drugs for Neurodegenerative Diseases. Mar. Drugs 2022, 20, 708. [Google Scholar] [CrossRef] [PubMed]
  17. Wattmo, C.; Wallin, Å.K. Early-versus Late-Onset Alzheimer’s Disease in Clinical Practice: Cognitive and Global Outcomes over 3 Years. Alzheimers Res. Ther. 2017, 9, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Finneran, D.J.; Nash, K.R. Neuroinflammation and fractalkine signaling in Alzheimer’s disease. J. Neuroinflamm. 2019, 16, 30. [Google Scholar] [CrossRef]
  19. Despres, C.; Byrne, C.; Qi, H.; Cantrelle, F.-X.; Huvent, I.; Chambraud, B.; Baulieu, E.-E.; Jacquot, Y.; Landrieu, I.; Lippens, G.; et al. Identification of the Tau phosphorylation pattern that drives its aggregation. Proc. Natl. Acad. Sci. USA 2017, 114, 9080–9085. [Google Scholar] [CrossRef] [Green Version]
  20. Liu, S.; Gao, J.; Zhu, M.; Liu, K.; Zhang, H.-L. Gut Microbiota and Dysbiosis in Alzheimer’s Disease: Implications for Pathogenesis and Treatment. Mol. Neurobiol. 2020, 57, 5026–5043. [Google Scholar] [CrossRef]
  21. Tian, Y.; Meng, L.; Zhang, Z. What is strain in neurodegenerative diseases? Cell. Mol. Life Sci. 2020, 77, 665–676. [Google Scholar] [CrossRef]
  22. Tiwari, S.; Atluri, V.; Kaushik, A.; Yndart, A.; Nair, M. Alzheimer’s disease: Pathogenesis, diagnostics, and therapeutics. Int. J. Nanomed. 2019, 14, 5541–5554. [Google Scholar] [CrossRef] [Green Version]
  23. Hardy, J.A.; Higgins, G.A. Alzheimer’s disease: The amyloid cascade hypothesis. Science 1992, 256, 184–185. [Google Scholar] [CrossRef]
  24. Silva, M.; Seijas, P.; Otero, P. Exploitation of Marine Molecules to Manage Alzheimer’s Disease. Mar. Drugs 2021, 19, 373. [Google Scholar] [CrossRef]
  25. Kinney, J.W.; BeMiller, S.M.; Murtishaw, A.S.; Leisgang, A.M.; Salazar, A.M.; Lamb, B.T. Inflammation as a central mechanism in Alzheimer’s disease. Alzheimers Dement. Transl. Res. Clin. Interv. 2018, 4, 575–590. [Google Scholar] [CrossRef] [PubMed]
  26. Pascoal, T.A.; Benedet, A.L.; Ashton, N.J.; Kang, M.S.; Therriault, J.; Chamoun, M.; Savard, M.; Lussier, F.Z.; Tissot, C.; Karikari, T.K.; et al. Microglial activation and tau propagate jointly across Braak stages. Nat. Med. 2021, 27, 1592–1599. [Google Scholar] [CrossRef] [PubMed]
  27. Zhong, W.; Wu, A.; Berglund, K.; Gu, X.; Jiang, M.Q.; Talati, J.; Zhao, J.; Wei, L.; Yu, S.P. Pathogenesis of sporadic Alzheimer’s disease by deficiency of NMDA receptor subunit GluN3A. Alzheimers Dement. 2021, 18, 222–239. [Google Scholar] [CrossRef]
  28. Naseri, N.N.; Wang, H.; Guo, J.; Sharma, M.; Luo, W. The complexity of tau in Alzheimer’s disease. Neurosci. Lett. 2019, 705, 183–194. [Google Scholar] [CrossRef] [PubMed]
  29. Van der Jeugd, A.; Hochgräfe, K.; Ahmed, T.; Decker, J.M.; Sydow, A.; Hofmann, A.; Wu, D.; Messing, L.; Balschun, D.; D’Hooge, R.; et al. Cognitive defects are reversible in inducible mice expressing pro-aggregant full-length human Tau. Acta Neuropathol. 2012, 123, 787–805. [Google Scholar] [CrossRef] [Green Version]
  30. Ferreira-Vieira, T.H.; Guimaraes, I.M.; Silva, F.R.; Ribeiro, F.M. Alzheimer’s disease: Targeting the Cholinergic System. Curr. Neuropharmacol. 2016, 14, 101–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Rosenberg, P.B.; Nowrangi, M.A.; Lyketsos, C.G. Neuropsychiatric symptoms in Alzheimer’s disease: What might be associated brain circuits? Mol. Asp. Med. 2015, 43–44, 25–37. [Google Scholar] [CrossRef] [Green Version]
  32. Martorana, A.; Koch, G. “Is Dopamine Involved in Alzheimer’s Disease?”. Front. Aging Neurosci. 2014, 6, 252. [Google Scholar] [CrossRef] [Green Version]
  33. Nitsch, R.M.; Deng, M.; Tennis, M.; Schoenfeld, D.; Growdon, J.H. The selective muscarinic M1 agonist AF102B decreases levels of total Abeta in cerebrospinal fluid of patients with Alzheimer’s disease. Ann. Neurol. 2000, 48, 913–918. [Google Scholar] [CrossRef]
  34. Jeon, J.; Dencker, D.; Wörtwein, G.; Woldbye, D.P.D.; Cui, Y.; Davis, A.A.; Levey, A.I.; Schütz, G.; Sager, T.N.; Mørk, A.; et al. A Subpopulation of Neuronal M4 Muscarinic Acetylcholine Receptors Plays a Critical Role in Modulating Dopamine-Dependent Behaviors. J. Neurosci. 2010, 30, 2396–2405. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, J.; Chang, L.; Song, Y.; Li, H.; Wu, Y. The Role of NMDA Receptors in Alzheimer’s Disease. Front. Neurosci. 2019, 13, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Role of Glutamate and NMDA Receptors in Alzheimer’s Disease—IOS Press. Available online: https://content.iospress.com/articles/journal-of-alzheimers-disease/jad160763 (accessed on 26 September 2022).
  37. Breijyeh, Z.; Karaman, R. Comprehensive Review on Alzheimer’s Disease: Causes and Treatment. Molecules 2020, 25, 5789. [Google Scholar] [CrossRef] [PubMed]
  38. Sharma, K. Cholinesterase Inhibitors as Alzheimer’s Therapeutics (Review). Mol. Med. Rep. 2019, 20, 1479–1487. [Google Scholar] [CrossRef] [Green Version]
  39. Green Valley Obtains IND Approval from U.S. FDA for Sodium Oligomannate’s International, Phase III Clinical Study—PR Newswire APAC. Available online: https://en.prnasia.com/story/278549-0.shtml (accessed on 3 November 2022).
  40. Carney, G.; Bassett, K.; Wright, J.M.; Maclure, M.; McGuire, N.; Dormuth, C.R. Comparison of cholinesterase inhibitor safety in real-world practice. Alzheimers Dement. Transl. Res. Clin. Interv. 2019, 5, 732–739. [Google Scholar] [CrossRef] [PubMed]
  41. Prince, M.; Bryce, R.; Albanese, E.; Wimo, A.; Ribeiro, W.; Ferri, C.P. The global prevalence of dementia: A systematic review and metaanalysis. Alzheimers Dement. 2013, 9, 63–75.e2. [Google Scholar] [CrossRef]
  42. Khoury, R.; Rajamanickam, J.; Grossberg, G.T. An update on the safety of current therapies for Alzheimer’s disease: Focus on rivastigmine. Ther. Adv. Drug Saf. 2018, 9, 171–178. [Google Scholar] [CrossRef] [PubMed]
  43. Syed, Y.Y. Sodium Oligomannate: First Approval. Drugs 2020, 80, 441–444. [Google Scholar] [CrossRef] [PubMed]
  44. Kumar, A.; Gupta, V.; Sharma, S. Donepezil. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  45. Birks, J.S.; Evans, J.G. Rivastigmine for Alzheimer’s disease. Cochrane Database Syst. Rev. 2015, 4, 1465–1858, CD001191. [Google Scholar] [CrossRef]
  46. Drugs to Treat Alzheimer’s Disease. J. Psychosoc. Nurs. Ment. Health Serv. 2018, 56, 5–6. [CrossRef]
  47. Léveillé, F.; El Gaamouch, F.; Gouix, E.; Lecocq, M.; Lobner, D.; Nicole, O.; Buisson, A. Neuronal viability is controlled by a functional relation between synaptic and extrasynaptic NMDA receptors. FASEB J. 2008, 22, 4258–4271. [Google Scholar] [CrossRef] [Green Version]
  48. Arndt, J.W.; Qian, F.; Smith, B.A.; Quan, C.; Kilambi, K.P.; Bush, M.W.; Walz, T.; Pepinsky, R.B.; Bussière, T.; Hamann, S.; et al. Structural and kinetic basis for the selectivity of aducanumab for aggregated forms of amyloid-β. Sci. Rep. 2018, 8, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Biogen. A Phase 3 Multicenter, Randomized, Double-Blind, Placebo-Controlled, Parallel-Group Study to Evaluate the Efficacy and Safety of Aducanumab (BIIB037) in Subjects With Early Alzheimer’s Disease. 2021; NCT02484547. Available online: https://clinicaltrials.gov/ct2/show/NCT02484547 (accessed on 9 August 2021).
  50. 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. Alzheimers Res. Ther. 2021, 13, 62. [Google Scholar] [CrossRef] [PubMed]
  51. Wang, X.; Sun, G.; Feng, T.; Zhang, J.; Huang, X.; Wang, T.; Geng, M. 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] [PubMed] [Green Version]
  52. Kong, L.N.; Geng, M.Y.; Mu, L.; Xin, X.L.; Yang, N.; Zuo, P.P. Effects of Acidic Oligose on Differentially Expressed Genes in the Mice Model of Alzheimer’s Disease by Microarray. Acta Pharm. Sin. 2005, 40, 1105–1109. Available online: https://en.cnki.com.cn/Article_en/CJFDTotal-YXXB200512009.htm (accessed on 20 August 2022).
  53. Jiang, R.W.; Du, X.G.; Zhang, X.; Wang, X.; Hu, D.Y.; Meng, T.; Chen, Y.L.; Geng, M.Y.; Shen, J.K. Synthesis and Bioassay of β-(1,4)-D-Mannans as Potential Agents against Alzheimer’s Disease. Acta Pharmacol. Sin. 2013, 34, 1585–1591. [Google Scholar] [CrossRef] [Green Version]
  54. Cummings, J.L.; Tong, G.; Ballard, C. Treatment Combinations for Alzheimer’s Disease: Current and Future Pharmacotherapy Options. J. Alzheimers Dis. 2019, 67, 779–794. [Google Scholar] [CrossRef] [Green Version]
  55. Amat-Ur-Rasool, H.; Ahmed, M.; Hasnain, S.; Carter, W. Anti-Cholinesterase Combination Drug Therapy as a Potential Treatment for Alzheimer’s Disease. Brain Sci. 2021, 11, 184. [Google Scholar] [CrossRef]
  56. Karthikeyan, A.; Joseph, A.; Nair, B.G. Promising bioactive compounds from the marine environment and their potential effects on various diseases. J. Genet. Eng. Biotechnol. 2022, 20, 14. [Google Scholar] [CrossRef]
  57. Zhang, L.; Hao, J.; Zheng, Y.; Su, R.; Liao, Y.; Gong, X.; Liu, L.; Wang, X. Fucoidan Protects Dopaminergic Neurons by Enhancing the Mitochondrial Function in a Rotenone-induced Rat Model of Parkinson’s Disease. Aging Dis. 2018, 9, 590–604. [Google Scholar] [CrossRef] [Green Version]
  58. Hao, C.; Wang, W.; Wang, S.; Zhang, L.; Guo, Y. An Overview of the Protective Effects of Chitosan and Acetylated Chitosan Oligosaccharides against Neuronal Disorders. Mar. Drugs 2017, 15, 89. [Google Scholar] [CrossRef] [Green Version]
  59. Liu, Q.; Xi, Y.; Wang, Q.; Liu, J.; Li, P.; Meng, X.; Liu, K.; Chen, W.; Liu, X.; Liu, Z. Mannan oligosaccharide attenuates cognitive and behavioral disorders in the 5xFAD Alzheimer’s disease mouse model via regulating the gut microbiota-brain axis. Brain Behav. Immun. 2021, 95, 330–343. [Google Scholar] [CrossRef] [PubMed]
  60. Yoon, J.-H.; Lee, N.; Youn, K.; Jo, M.; Kim, H.-R.; Lee, D.-S.; Ho, C.-T.; Jun, M. Dieckol Ameliorates Aβ Production via PI3K/Akt/GSK-3β Regulated APP Processing in SweAPP N2a Cell. Mar. Drugs 2021, 19, 152. [Google Scholar] [CrossRef] [PubMed]
  61. Mouzat, K.; Chudinova, A.; Polge, A.; Kantar, J.; Camu, W.; Raoul, C.; Lumbroso, S. Regulation of Brain Cholesterol: What Role Do Liver X Receptors Play in Neurodegenerative Diseases? Int. J. Mol. Sci. 2019, 20, 3858. [Google Scholar] [CrossRef]
  62. Hannan, A.; Dash, R.; Sohag, A.A.M.; Moon, I.S. Deciphering Molecular Mechanism of the Neuropharmacological Action of Fucosterol through Integrated System Pharmacology and In Silico Analysis. Mar. Drugs 2019, 17, 639. [Google Scholar] [CrossRef] [Green Version]
  63. Torregrosa-Crespo, J.; Montero, Z.; Fuentes, J.L.; García-Galbis, M.R.; Garbayo, I.; Vílchez, C.; Martínez-Espinosa, R.M. Exploring the Valuable Carotenoids for the Large-Scale Production by Marine Microorganisms. Mar. Drugs 2018, 16, 203. [Google Scholar] [CrossRef] [Green Version]
  64. Sathasivam, R.; Ki, J.-S. A Review of the Biological Activities of Microalgal Carotenoids and Their Potential Use in Healthcare and Cosmetic Industries. Mar. Drugs 2018, 16, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wang, J.; Li, L.; Wang, Z.; Cui, Y.; Tan, X.; Yuan, T.; Liu, Q.; Liu, Z.; Liu, X. Supplementation of lycopene attenuates lipopolysaccharide-induced amyloidogenesis and cognitive impairments via mediating neuroinflammation and oxidative stress. J. Nutr. Biochem. 2018, 56, 16–25. [Google Scholar] [CrossRef]
  66. Gegunde, S.; Alfonso, A.; Alonso, E.; Alvariño, R.; Botana, L.M. Gracilin-Derivatives as Lead Compounds for Anti-inflammatory Effects. Cell. Mol. Neurobiol. 2020, 40, 603–615. [Google Scholar] [CrossRef]
  67. Palem, J.R.; Mudit, M.; Hsia, S.-C.V.; El Sayed, K.A. Discovery and preliminary structure-activity relationship of the marine natural product manzamines as herpes simplex virus type-1 inhibitors. Z. Nat. C 2017, 72, 49–54. [Google Scholar] [CrossRef]
  68. Hu, Y.; Chen, S.; Yang, F.; Dong, S. Marine Indole Alkaloids—Isolation, Structure and Bioactivities. Mar. Drugs 2021, 19, 658. [Google Scholar] [CrossRef]
  69. Caltagirone, C.; Ferrannini, L.; Marchionni, N.; Nappi, G.; Scapagnini, G.; Trabucchi, M. The potential protective effect of tramiprosate (homotaurine) against Alzheimer’s disease: A review. Aging Clin. Exp. Res. 2012, 24, 580–587. [Google Scholar] [CrossRef] [PubMed]
  70. Emartorana, A.; Lorenzo, F.E.; Manenti, G.; Esemprini, R.; Koch, G. Homotaurine Induces Measurable Changes of Short Latency Afferent Inhibition in a Group of Mild Cognitive Impairment Individuals. Front. Aging Neurosci. 2014, 6, 254. [Google Scholar] [CrossRef]
  71. Otero, P.; Alfonso, A.; Vieytes, M.R.; Cabado, A.G.; Vieites, J.M.; Botana, L.M. Effects of environmental regimens on the toxin profile ofAlexandrium ostenfeldii. Environ. Toxicol. Chem. 2010, 29, 301–310. [Google Scholar] [CrossRef]
  72. Russo, P.; Kisialiou, A.; Lamonaca, P.; Moroni, R.; Prinzi, G.; Fini, M. New Drugs from Marine Organisms in Alzheimer’s Disease. Mar. Drugs 2015, 14, 5. [Google Scholar] [CrossRef] [Green Version]
  73. Natural Polysaccharide: Modification and Application. Available online: http://qikan.cqvip.com/Qikan/Article/Detail?id=7104613483 (accessed on 26 September 2022).
  74. Apostolova, E.; Lukova, P.; Baldzhieva, A.; Katsarov, P.; Nikolova, M.; Iliev, I.; Peychev, L.; Trica, B.; Oancea, F.; Delattre, C.; et al. Immunomodulatory and Anti-Inflammatory Effects of Fucoidan: A Review. Polymers 2020, 12, 2338. [Google Scholar] [CrossRef]
  75. Wang, Y.; Xing, M.; Cao, Q.; Ji, A.; Liang, H.; Song, S. Biological Activities of Fucoidan and the Factors Mediating Its Therapeutic Effects: A Review of Recent Studies. Mar. Drugs 2019, 17, 183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Aleissa, M.S.; Alkahtani, S.; Eldaim, M.A.A.; Ahmed, A.M.; Bungău, S.G.; Almutairi, B.; Bin-Jumah, M.; AlKahtane, A.A.; Alyousif, M.S.; Abdel-Daim, M.M. Fucoidan Ameliorates Oxidative Stress, Inflammation, DNA Damage, and Hepatorenal Injuries in Diabetic Rats Intoxicated with Aflatoxin B1. Oxidative Med. Cell. Longev. 2020, 2020, 9316751. [Google Scholar] [CrossRef] [Green Version]
  77. Wu, H.; Gao, S.; Terakawa, S. Inhibitory effects of fucoidan on NMDA receptors and l-type Ca2+ channels regulating the Ca2+ responses in rat neurons. Pharm. Biol. 2019, 57, 1–7. [Google Scholar] [CrossRef] [Green Version]
  78. Hsieh, C.-H.; Lu, C.-H.; Kuo, Y.-Y.; Lin, G.-B.; Chao, C.-Y. The protective effect of non-invasive low intensity pulsed electric field and fucoidan in preventing oxidative stress-induced motor neuron death via ROCK/Akt pathway. PLoS ONE 2019, 14, e0214100. [Google Scholar] [CrossRef]
  79. Ji, H.L.; Park, J.E.; Ji, S.H. Fucoidan Stimulates Glucose Uptake via the PI3K/AMPK Pathway and Increases Insulin Sensitivity in 3T3-L1 Adipocytes. Korean Soc. Life Sci. 2021, 31, 1–9. [Google Scholar] [CrossRef]
  80. Wei, H.; Gao, Z.; Zheng, L.; Zhang, C.; Liu, Z.; Yang, Y.; Teng, H.; Hou, L.; Yin, Y.; Zou, X. Protective Effects of Fucoidan on Aβ25–35 and d-Gal-Induced Neurotoxicity in PC12 Cells and d-Gal-Induced Cognitive Dysfunction in Mice. Mar. Drugs 2017, 15, 77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Liu, H.; Wang, J.; Zhang, Q.; Zhang, H. The effect of different substitute groups and molecular weights of fucoidan on neuroprotective and anticomplement activity. Int. J. Biol. Macromol. 2018, 113, 82–89. [Google Scholar] [CrossRef] [PubMed]
  82. Tsvetkov, Y.E.; Paulovičová, E.; Paulovičová, L.; Farkaš, P.; Nifantiev, N.E. Synthesis of Biotin-Tagged Chitosan Oligosaccharides and Assessment of Their Immunomodulatory Activity. Front. Chem. 2020, 8, 554732. [Google Scholar] [CrossRef] [PubMed]
  83. Kumar, M.N.R. A Review of Chitin and Chitosan Applications. React. Funct. Polym. 2000, 46, 1–27. [Google Scholar] [CrossRef]
  84. Jeon, Y.-J.; Park, P.-J.; Kim, S.-K. Antimicrobial Effect of Chitooligosaccharides Produced by Bioreactor. Carbohydr. Polym. 2001, 44, 71–76. [Google Scholar] [CrossRef]
  85. Liaqat, F.; Eltem, R. Chitooligosaccharides and their biological activities: A comprehensive review. Carbohydr. Polym. 2018, 184, 243–259. [Google Scholar] [CrossRef]
  86. Khodagholi, F.; Eftekharzadeh, B.; Maghsoudi, N.; Rezaei, P.F. Retracted Article: Chitosan prevents oxidative stress-induced amyloid β formation and cytotoxicity in NT2 neurons: Involvement of transcription factors Nrf2 and NF-κB. Mol. Cell. Biochem. 2009, 337, 39–51. [Google Scholar] [CrossRef]
  87. Yoon, N.Y.; Ngo, D.-N.; Kim, S.-K. Acetylcholinesterase inhibitory activity of novel chitooligosaccharide derivatives. Carbohydr. Polym. 2009, 78, 869–872. [Google Scholar] [CrossRef]
  88. Zhu, L.; Li, R.; Jiao, S.; Wei, J.; Yan, Y.; Wang, Z.A.; Li, J.; Du, Y. Blood-Brain Barrier Permeable Chitosan Oligosaccharides Interfere with β-amyloid Aggregation and Alleviate β-Amyloid Protein Mediated Neurotoxicity and Neuroinflammation in a Dose- and Degree of Polymerization-Dependent Manner. Mar. Drugs 2020, 18, 488. [Google Scholar] [CrossRef]
  89. Xue, Y.; Wang, N.; Zeng, Z.; Huang, J.; Xiang, Z.; Guan, Y.-Q. Neuroprotective effect of chitosan nanoparticle gene delivery system grafted with acteoside (ACT) in Parkinson’s disease models. J. Mater. Sci. Technol. 2020, 43, 197–207. [Google Scholar] [CrossRef]
  90. Faustino, M.; Durão, J.; Pereira, C.F.; Pintado, M.E.; Carvalho, A.P. Mannans and mannan oligosaccharides (MOS) from Saccharomyces cerevisiae—A sustainable source of functional ingredients. Carbohydr. Polym. 2021, 272, 118467. [Google Scholar] [CrossRef] [PubMed]
  91. Singh, S.; Singh, G.; Arya, S.K. Mannans: An overview of properties and application in food products. Int. J. Biol. Macromol. 2018, 119, 79–95. [Google Scholar] [CrossRef] [PubMed]
  92. Crispi, S.; Filosa, S. Novel perspectives for neurodegeneration prevention: Effects of bioactive polyphenols. Neural Regen. Res. 2021, 16, 1411–1412. [Google Scholar] [CrossRef] [PubMed]
  93. Li, C.; Meng, P.; Zhang, B.; Kang, H.; Wen, H.; Schluesener, H.; Cao, Z.; Zhang, Z. Computer-aided identification of protein targets of four polyphenols in Alzheimer’s disease (AD) and validation in a mouse AD model. J. Biomed. Res. 2019, 33, 101–112. [Google Scholar] [CrossRef] [PubMed]
  94. Sang, V.T.; Hung, N.D.; Se-Kwon, K. Pharmaceutical properties of marine polyphenols: An overview. Acta Pharm. Sci. 2019, 57, 217. [Google Scholar] [CrossRef] [Green Version]
  95. Ying, S.-W.; Futter, M.; Rosenblum, K.; Webber, M.J.; Hunt, S.P.; Bliss, T.V.P.; Bramham, C.R. Brain-Derived Neurotrophic Factor Induces Long-Term Potentiation in Intact Adult Hippocampus: Requirement for ERK Activation Coupled to CREB and Upregulation of Arc Synthesis. J. Neurosci. 2002, 22, 1532–1540. [Google Scholar] [CrossRef] [Green Version]
  96. Um, M.Y.; Lim, D.W.; Son, H.J.; Cho, S.; Lee, C. Phlorotannin-rich fraction from Ishige foliacea brown seaweed prevents the scopolamine-induced memory impairment via regulation of ERK-CREB-BDNF pathway. J. Funct. Foods 2018, 40, 110–116. [Google Scholar] [CrossRef]
  97. Zambrano, P.; Suwalsky, M.; Jemiola-Rzeminska, M.; Strzalka, K.; Sepúlveda, B.; Gallardo, M.J.; Aguilar, L.F. The acetylcholinesterase (AChE) inhibitor and anti-Alzheimer drug donepezil interacts with human erythrocytes. Biochim. Biophys. Acta Biomembr. 2019, 1861, 1078–1085. [Google Scholar] [CrossRef]
  98. Hannan, A.; Dash, R.; Haque, N.; Mohibbullah; Sohag, A.; Rahman, A.; Uddin, J.; Alam, M.; Moon, I.S. Neuroprotective Potentials of Marine Algae and Their Bioactive Metabolites: Pharmacological Insights and Therapeutic Advances. Mar. Drugs 2020, 10, 347. [Google Scholar] [CrossRef]
  99. Minter, M.R.; Taylor, J.M.; Crack, P.J. The contribution of neuroinflammation to amyloid toxicity in Alzheimer’s disease. J. Neurochem. 2016, 136, 457–474. [Google Scholar] [CrossRef] [Green Version]
  100. Rahman, A.; Dash, R.; Sohag, A.; Alam, M.; Rhim, H.; Ha, H.; Moon, I.; Uddin, J.; Hannan, A. Prospects of Marine Sterols against Pathobiology of Alzheimer’s Disease: Pharmacological Insights and Technological Advances. Mar. Drugs 2021, 19, 167. [Google Scholar] [CrossRef] [PubMed]
  101. Zhou, X.; Lu, Y.; Lin, X.; Yang, B.; Yang, X.; Liu, Y. Brominated aliphatic hydrocarbons and sterols from the sponge Xestospongia testudinaria with their bioactivities. Chem. Phys. Lipids 2011, 164, 703–706. [Google Scholar] [CrossRef] [PubMed]
  102. 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]
  103. 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] [PubMed]
  104. Olasehinde, T.A.; Olaniran, A.O.; Okoh, A.I. Macroalgae as a Valuable Source of Naturally Occurring Bioactive Compounds for the Treatment of Alzheimer’s Disease. Mar. Drugs 2019, 17, 609. [Google Scholar] [CrossRef] [Green Version]
  105. Galasso, C.; Corinaldesi, C.; Sansone, C. Carotenoids from Marine Organisms: Biological Functions and Industrial Applications. Antioxidants 2017, 6, 96. [Google Scholar] [CrossRef] [Green Version]
  106. 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] [Green Version]
  107. 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]
  108. Manochkumar, J.; Doss, C.G.P.; El-Seedi, H.R.; Efferth, T.; Ramamoorthy, S. The neuroprotective potential of carotenoids in vitro and in vivo. Phytomedicine 2021, 91, 153676. [Google Scholar] [CrossRef]
  109. Ohgami, K.; Shiratori, K.; Kotake, S.; Nishida, T.; Mizuki, N.; Yazawa, K.; Ohno, S. Effects of Astaxanthin on Lipopolysaccharide-Induced Inflammation In Vitro and In Vivo. Investig. Opthalmol. Vis. Sci. 2003, 44, 2694–2701. [Google Scholar] [CrossRef] [Green Version]
  110. Bonito, M.C.; Cicala, C.; Marcotullio, M.C.; Maione, F.; Mascolo, N. Biological Activity of Bicyclic and Tricyclic Diterpenoids from Salvia Species of Immediate Pharmacological and Pharmaceutical Interest. Nat. Prod. Commun. 2011, 6, 1205–1215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Leirós, M.; Alonso, E.; Rateb, M.E.; Houssen, W.E.; Ebel, R.; Jaspars, M.; Alfonso, A.; Botana, L.M. Gracilins: Spongionella-derived promising compounds for Alzheimer disease. Neuropharmacology 2015, 93, 285–293. [Google Scholar] [CrossRef] [PubMed]
  112. Choi, D.-Y.; Choi, H. Natural products from marine organisms with neuroprotective activity in the experimental models of Alzheimer’s disease, Parkinson’s disease and ischemic brain stroke: Their molecular targets and action mechanisms. Arch. Pharmacal Res. 2015, 38, 139–170. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, Y.; Wang, Q.; Zheng, D.-K.; Zhang, D.; Xie, Z.; Hu, J.-W.; Xie, X.-H.; Li, J.; Jiang, S.-P. Abietane diterpenoids with neuroprotective activities from Phlegmariurus carinatus. Nat. Prod. Res. 2022, 36, 6006–6011. [Google Scholar] [CrossRef]
  114. Abbasov, M.E.; Alvariño, R.; Chaheine, C.M.; Alonso, E.; Sánchez, J.A.; Conner, M.L.; Alfonso, A.; Jaspars, M.; Botana, L.M.; Romo, D. Simplified immunosuppressive and neuroprotective agents based on gracilin A. Nat. Chem. 2019, 11, 342–350. [Google Scholar] [CrossRef]
  115. Alvariño, R.; Alonso, E.; Abbasov, M.E.; Chaheine, C.M.; Conner, M.L.; Romo, D.; Alfonso, A.; Botana, L.M. Gracilin A Derivatives Target Early Events in Alzheimer’s Disease: In Vitro Effects on Neuroinflammation and Oxidative Stress. ACS Chem. Neurosci. 2019, 10, 4102–4111. [Google Scholar] [CrossRef]
  116. Kabir, T.; Uddin, S.; Jeandet, P.; Emran, T.; Mitra, S.; Albadrani, G.; Sayed, A.; Abdel-Daim, M.; Simal-Gandara, J. Anti-Alzheimer’s Molecules Derived from Marine Life: Understanding Molecular Mechanisms and Therapeutic Potential. Mar. Drugs 2021, 19, 251. [Google Scholar] [CrossRef]
  117. Li, H.; Liu, X.; Li, X.; Hu, Z.; Wang, L. Novel Harziane Diterpenes from Deep-Sea Sediment Fungus Trichoderma sp. SCSIOW21 and Their Potential Anti-Inflammatory Effects. Mar. Drugs 2021, 19, 689. [Google Scholar] [CrossRef]
  118. Althagbi, H.I.; Alarif, W.M.; Al-Footy, K.O.; Abdel-Lateff, A. Marine-Derived Macrocyclic Alkaloids (MDMAs): Chemical and Biological Diversity. Mar. Drugs 2020, 18, 368. [Google Scholar] [CrossRef]
  119. Karan, D.; Dubey, S.; Pirisi, L.; Nagel, A.; Pina, I.; Choo, Y.-M.; Hamann, M.T. The Marine Natural Product Manzamine A Inhibits Cervical Cancer by Targeting the SIX1 Protein. J. Nat. Prod. 2020, 83, 286–295. [Google Scholar] [CrossRef]
  120. Yousaf, M.; Hammond, N.L.; Peng, J.; Wahyuono, S.; McIntosh, K.A.; Charman, W.N.; Mayer, A.A.M.S.; Hamann, M.T. New Manzamine Alkaloids from an Indo-Pacific Sponge. Pharmacokinetics, Oral Availability, and the Significant Activity of Several Manzamines against HIV-I, AIDS Opportunistic Infections, and Inflammatory Diseases. J. Med. Chem. 2004, 47, 3512–3517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Rao, K.V.; Donia, M.S.; Peng, J.; Garcia-Palomero, E.; Alonso, D.; Martinez, A.; Medina, M.; Franzblau, S.G.; Tekwani, B.L.; Khan, S.I.; et al. Manzamine B and E and Ircinal A Related Alkaloids from an Indonesian Acanthostrongylophora Sponge and Their Activity against Infectious, Tropical Parasitic, and Alzheimer’s Diseases. J. Nat. Prod. 2006, 69, 1034–1040. [Google Scholar] [CrossRef] [Green Version]
  122. Lima, E.; Medeiros, J. Marine Organisms as Alkaloid Biosynthesizers of Potential Anti-Alzheimer Agents. Mar. Drugs 2022, 20, 75. [Google Scholar] [CrossRef] [PubMed]
  123. Miguel-Gordo, M.; Gegunde, S.; Calabro, K.; Jennings, L.K.; Alfonso, A.; Genta-Jouve, G.; Vacelet, J.; Botana, L.M.; Thomas, O.P. Bromotryptamine and Bromotyramine Derivatives from the Tropical Southwestern Pacific Sponge Narrabeena nigra. Mar. Drugs 2019, 17, 319. [Google Scholar] [CrossRef] [Green Version]
  124. Zhang, F.; Shi, J.-S.; Li, D.-D.; Zheng, C.-Q. Potential neuroprotection by Dendrobium nobile Lindl alkaloid in Alzheimer’s disease models. Neural Regen. Res. 2022, 17, 972–977. [Google Scholar] [CrossRef] [PubMed]
  125. Aisen, P.S.; Saumier, D.; Briand, R.; Laurin, J.; Gervais, F.; Tremblay, P.; Garceau, D. A Phase II study targeting amyloid-with 3APS in mild-to-moderate Alzheimer disease. Neurology 2006, 67, 1757–1763. [Google Scholar] [CrossRef] [PubMed]
  126. Jakaria, M.; Azam, S.; Haque, M.E.; Jo, S.-H.; Uddin, M.S.; Kim, I.-S.; Choi, D.-K. Taurine and its analogs in neurological disorders: Focus on therapeutic potential and molecular mechanisms. Redox Biol. 2019, 24, 101223. [Google Scholar] [CrossRef]
  127. Shivaraj, M.C.; Marcy, G.; Low, G.; Ryu, J.R.; Zhao, X.; Rosales, F.J.; Goh, E.L.K. Taurine Induces Proliferation of Neural Stem Cells and Synapse Development in the Developing Mouse Brain. PLoS ONE 2012, 7, e42935. [Google Scholar] [CrossRef]
  128. Mehdinia, A.; Fumani, N.S.; Kayyal, T.B.; Ghaderiardakani, F. Homotaurine of Marine Macroalgae of the Persian Gulf as a Potential Treatment Agent for Alzheimer. Mar. Sci. 2018, 9, 1–8. [Google Scholar]
  129. Chekan, J.R.; Fallon, T.R.; Moore, B.S. Biosynthesis of marine toxins. Curr. Opin. Chem. Biol. 2020, 59, 119–129. [Google Scholar] [CrossRef]
  130. Kem, W.R.; Andrud, K.; Bruno, G.; Xing, H.; Soti, F.; Talley, T.T.; Taylor, P. Interactions of Nereistoxin and Its Analogs with Vertebrate Nicotinic Acetylcholine Receptors and Molluscan ACh Binding Proteins. Mar. Drugs 2022, 20, 49. [Google Scholar] [CrossRef]
  131. Touzet, N.; Franco, J.M.; Raine, R. Morphogenetic diversity and biotoxin composition of Alexandrium (Dinophyceae) in Irish coastal waters. Harmful Algae 2008, 7, 782–797. [Google Scholar] [CrossRef]
  132. Hu, T.; Burton, I.W.; Cembella, A.D.; Curtis, J.M.; Quilliam, M.A.; Walter, J.A.; Wright, J.L.C. Characterization of Spirolides A, C, and 13-Desmethyl C, New Marine Toxins Isolated from Toxic Plankton and Contaminated Shellfish. J. Nat. Prod. 2001, 64, 308–312. [Google Scholar] [CrossRef]
  133. Boente-Juncal, A.; Méndez, A.G.; Vale, C.; Vieytes, M.R.; Botana, L.M. In Vitro Effects of Chronic Spirolide Treatment on Human Neuronal Stem Cell Differentiation and Cholinergic System Development. ACS Chem. Neurosci. 2018, 9, 1441–1452. [Google Scholar] [CrossRef]
  134. Alonso, E.; Vale, C.; Vieytes, M.R.; Laferla, F.M.; Giménez-Llort, L.; Botana, L.M. 13-Desmethyl spirolide-C is neuroprotective and reduces intracellular Aβ and hyperphosphorylated tau in vitro. Neurochem. Int. 2011, 59, 1056–1065. [Google Scholar] [CrossRef]
  135. Botelho, M.J.; Milinovic, J.; Bandarra, N.M.; Vale, C. Alzheimer’s Disease and Toxins Produced by Marine Dinoflagellates: An Issue to Explore. Mar. Drugs 2022, 20, 253. [Google Scholar] [CrossRef]
  136. Mahmood, N.A.; Carmichael, W.W. Anatoxin-a(s), an anticholinesterase from the cyanobacterium Anabaena flos-aquae NRC-525-17. Toxicon 1987, 25, 1221–1227. [Google Scholar] [CrossRef]
  137. Castaneda, A.; Ferraz, R.; Vieira, M.; Cardoso, I.; Vasconcelos, V.; Martins, R. Bridging Cyanobacteria to Neurodegenerative Diseases: A New Potential Source of Bioactive Compounds against Alzheimer’s Disease. Mar. Drugs 2021, 19, 343. [Google Scholar] [CrossRef]
  138. Aldholmi, M.; Marchand, P.; Ourliac-Garnier, I.; Le Pape, P.; Ganesan, A. A Decade of Antifungal Leads from Natural Products: 2010–2019. Pharmaceuticals 2019, 12, 182. [Google Scholar] [CrossRef] [Green Version]
  139. Alvariño, R.; Alonso, E.; Lacret, R.; Oves-Costales, D.; Genilloud, O.; Reyes, F.; Alfonso, A.; Botana, L.M. Caniferolide A, a Macrolide from Streptomyces caniferus, Attenuates Neuroinflammation, Oxidative Stress, Amyloid-Beta, and Tau Pathology in Vitro. Mol. Pharm. 2019, 16, 1456–1466. [Google Scholar] [CrossRef]
  140. Sun, M.-K.; Nelson, T.J.; Alkon, D.L. Towards universal therapeutics for memory disorders. Trends Pharmacol. Sci. 2015, 36, 384–394. [Google Scholar] [CrossRef]
  141. Hongpaisan, J.; Sun, M.-K.; Alkon, D.L. PKC Activation Prevents Synaptic Loss, A Elevation, and Cognitive Deficits in Alzheimer’s Disease Transgenic Mice. J. Neurosci. 2011, 31, 630–643. [Google Scholar] [CrossRef] [Green Version]
  142. Neurotrope Announces Positive Top-Line Results From Its Phase 2a Study of Bryostatin-1 in Alzheimer’s Disease. Available online: https://www.prnewswire.com/news-releases/neurotrope-announces-positive-top-line-results-from-its-phase-2a-study-of-bryostatin-1-in-alzheimers-disease-300038367.html (accessed on 3 November 2022).
  143. Farlow, M.R.; Thompson, R.E.; Wei, L.-J.; Tuchman, A.J.; Grenier, E.; Crockford, D.; Wilke, S.; Benison, J.; Alkon, D.L. A Randomized, Double-Blind, Placebo-Controlled, Phase II Study Assessing Safety, Tolerability, and Efficacy of Bryostatin in the Treatment of Moderately Severe to Severe Alzheimer’s Disease. J. Alzheimers Dis. 2019, 67, 555–570. [Google Scholar] [CrossRef] [Green Version]
  144. Weidner, W.S.; Barbarino, P. P4–443: The State Of The Art Of Dementia Research: New Frontiers. Alzheimers Dement. 2019, 15, P1473. [Google Scholar] [CrossRef]
  145. Cacabelos, R. How plausible is an Alzheimer’s disease vaccine? Expert Opin. Drug Discov. 2019, 15, 1–6. [Google Scholar] [CrossRef] [Green Version]
  146. Reich, N.; Hölscher, C. Acylated Ghrelin as a Multi-Targeted Therapy for Alzheimer’s and Parkinson’s Disease. Front. Neurosci. 2020, 14, 614828. [Google Scholar] [CrossRef]
  147. Cabrera-Pardo, J.R.; Fuentealba, J.; Gavilán, J.; Cajas, D.; Becerra, J.; Napiórkowska, M. Exploring the Multi–Target Neuroprotective Chemical Space of Benzofuran Scaffolds: A New Strategy in Drug Development for Alzheimer’s Disease. Front. Pharmacol. 2020, 10, 1679. [Google Scholar] [CrossRef]
  148. Bolea, I.; Gella, A.; Unzeta, M. Propargylamine-derived multitarget-directed ligands: Fighting Alzheimer’s disease with monoamine oxidase inhibitors. J. Neural Transm. 2013, 120, 893–902. [Google Scholar] [CrossRef]
  149. Xuan, Z.; Gu, X.; Yan, S.; Xie, Y.; Zhou, Y.; Zhang, H.; Jin, H.; Hu, S.; Mak, M.S.H.; Zhou, D.; et al. Dimeric Tacrine(10)-hupyridone as a Multitarget-Directed Ligand To Treat Alzheimer’s Disease. ACS Chem. Neurosci. 2021, 12, 2462–2477. [Google Scholar] [CrossRef]
  150. Anisha, G.S.; Padmakumari, S.; Patel, A.K.; Pandey, A.; Singhania, R.R. Fucoidan from Marine Macroalgae: Biological Actions and Applications in Regenerative Medicine, Drug Delivery Systems and Food Industry. Bioengineering 2022, 9, 472. [Google Scholar] [CrossRef]
  151. Vargas, M.; Gonzalez-Martinez, C. Recent Patents on Food Applications of Chitosan. Recent Pat. Food Nutr. Agric. 2010, 2, 121–128. [Google Scholar] [CrossRef]
  152. Wang, S.-H.; Huang, C.-Y.; Chen, C.-Y.; Chang, C.-C.; Huang, C.-Y.; Dong, C.-D.; Chang, J.-S. Isolation and purification of brown algae fucoidan from Sargassum siliquosum and the analysis of anti-lipogenesis activity. Biochem. Eng. J. 2021, 165, 107798. [Google Scholar] [CrossRef]
  153. Lourenço-Lopes, C.; Garcia-Oliveira, P.; Carpena, M.; Fraga-Corral, M.; Jimenez-Lopez, C.; Pereira, A.G.; Prieto, M.A.; Simal-Gandara, J. Scientific Approaches on Extraction, Purification and Stability for the Commercialization of Fucoxanthin Recovered from Brown Algae. Foods 2020, 9, 1113. [Google Scholar] [CrossRef]
  154. Priyadarsani, S.; Patel, A.S.; Kar, A.; Dash, S. Process optimization for the supercritical carbondioxide extraction of lycopene from ripe grapefruit (Citrus paradisi) endocarp. Sci. Rep. 2021, 11, 10273. [Google Scholar] [CrossRef]
  155. Obluchinskaya, E.; Pozharitskaya, O.; Zakharova, L.; Daurtseva, A.; Flisyuk, E.; Shikov, A. Efficacy of Natural Deep Eutectic Solvents for Extraction of Hydrophilic and Lipophilic Compounds from Fucus vesiculosus. Molecules 2021, 26, 4198. [Google Scholar] [CrossRef]
  156. Trombino, F.C.A.C.R.S.; Curcio, F.; Cassano, R. Nano- and Micro-Technologies Applied to Food Nutritional Ingredients. Curr. Drug Deliv. 2021, 18, 670–678. [Google Scholar] [CrossRef]
  157. ElKhoury, K.; Sanchez-Gonzalez, L.; Lavrador, P.; Almeida, R.; Gaspar, V.; Kahn, C.; Cleymand, F.; Arab-Tehrany, E.; Mano, J.F. Gelatin Methacryloyl (GelMA) Nanocomposite Hydrogels Embedding Bioactive Naringin Liposomes. Polymers 2020, 12, 2944. [Google Scholar] [CrossRef]
  158. Zhang, D.; Zhong, D.; Ouyang, J.; He, J.; Qi, Y.; Chen, W.; Zhang, X.; Tao, W.; Zhou, M. Microalgae-based oral microcarriers for gut microbiota homeostasis and intestinal protection in cancer radiotherapy. Nat. Commun. 2022, 13, 1413. [Google Scholar] [CrossRef]
  159. Zhong, D.; Zhang, D.; Chen, W.; He, J.; Ren, C.; Zhang, X.; Kong, N.; Tao, W.; Zhou, M. Orally deliverable strategy based on microalgal biomass for intestinal disease treatment. Sci. Adv. 2021, 7, eabi9265. [Google Scholar] [CrossRef]
  160. Jeromin, A.; Bowser, R. Biomarkers in Neurodegenerative Diseases. Nat. Med. 2017, 15, 491–528. [Google Scholar] [CrossRef]
  161. Ehrenberg, A.J.; Khatun, A.; Coomans, E.; Betts, M.J.; Capraro, F.; Thijssen, E.H.; Senkevich, K.; Bharucha, T.; Jafarpour, M.; Young, P.N.E.; et al. Relevance of biomarkers across different neurodegenerative diseases. Alzheimers Res. Ther. 2020, 12, 56. [Google Scholar] [CrossRef] [PubMed]
  162. Ferreira, I.; Rauter, A.P.; Bandarra, N.M. Marine Sources of DHA-Rich Phospholipids with Anti-Alzheimer Effect. Mar. Drugs 2022, 20, 662. [Google Scholar] [CrossRef] [PubMed]
  163. da Silva, A.P.G. Fighting coronaviruses with natural polyphenols. Biocatal. Agric. Biotechnol. 2021, 37, 102179. [Google Scholar] [CrossRef]
Figure 1. Key factors in the pathogenesis of AD. Aβ, amyloid-beta; AD, Alzheimer’s disease; APP, amyloid precursor protein; IR, insulin resistance; NFT, neurofibrillary tangle; OS, oxidative stress; ROS, reactive oxygen species; TBI, traumatic brain injury. (Re-produced with permission from [20], Springer Nature, 2020.)
Figure 1. Key factors in the pathogenesis of AD. Aβ, amyloid-beta; AD, Alzheimer’s disease; APP, amyloid precursor protein; IR, insulin resistance; NFT, neurofibrillary tangle; OS, oxidative stress; ROS, reactive oxygen species; TBI, traumatic brain injury. (Re-produced with permission from [20], Springer Nature, 2020.)
Marinedrugs 21 00043 g001
Figure 2. The structure of AduFab with bound Aβ1-11 peptide. (a) Cartoon representation of AduFab showing heavy chain in green, light chain in cyan, and Aβ1-11 peptide in magenta, with nitrogen and oxygen atoms displayed in blue and red, respectively. L1 to L3 and H1 to H3 indicate the CDRs in the light and heavy chain, respectively. (b) Detailed view of the binding interface between AduFab and the Aβ1-11 peptide, with key interface residues of AduFab within 4 Å of the Aβ peptide shown and labeled. An omit electron density map contoured at 3.0 σ is shown as mesh and superposed on the Aβ peptide. (Re-produced with permission from [48], Springer Nature, 2018.)
Figure 2. The structure of AduFab with bound Aβ1-11 peptide. (a) Cartoon representation of AduFab showing heavy chain in green, light chain in cyan, and Aβ1-11 peptide in magenta, with nitrogen and oxygen atoms displayed in blue and red, respectively. L1 to L3 and H1 to H3 indicate the CDRs in the light and heavy chain, respectively. (b) Detailed view of the binding interface between AduFab and the Aβ1-11 peptide, with key interface residues of AduFab within 4 Å of the Aβ peptide shown and labeled. An omit electron density map contoured at 3.0 σ is shown as mesh and superposed on the Aβ peptide. (Re-produced with permission from [48], Springer Nature, 2018.)
Marinedrugs 21 00043 g002
Figure 3. Chemical structures of marine molecules with neuroprotective effects against Alzheimer’s Disease: fucoidan (A), chitosan (B), mannan (C), dieckol (D), fucosterol (E), astaxanthin (F), lycopene (G), gracilins A (H), manzamine A (I), indole (J), homotaurine (K), 13-desMeC (L), caniferolide A (M) and bryostatin 1 (N).
Figure 3. Chemical structures of marine molecules with neuroprotective effects against Alzheimer’s Disease: fucoidan (A), chitosan (B), mannan (C), dieckol (D), fucosterol (E), astaxanthin (F), lycopene (G), gracilins A (H), manzamine A (I), indole (J), homotaurine (K), 13-desMeC (L), caniferolide A (M) and bryostatin 1 (N).
Marinedrugs 21 00043 g003aMarinedrugs 21 00043 g003b
Table 1. Several related information about drugs currently used for AD treatment.
Table 1. Several related information about drugs currently used for AD treatment.
Drug’s NameChemical StructuresPharmacological MechanismRef.
TacrineMarinedrugs 21 00043 i001Inhibition of cholinesterase (both AChE and BChE)[40]
Donepezile HClMarinedrugs 21 00043 i002Inhibition of AChE[41]
RivastigmineMarinedrugs 21 00043 i003Reversible double inhibitors of AChE and BChE[42]
GalantamineMarinedrugs 21 00043 i004Weak competitive and reversible cholinesterase inhibitors[38]
MemantineMarinedrugs 21 00043 i005Antagonist of the NMDAR[37]
sodium oligomannaten = 1~9; m = 0, 1, 2; m’ = 0,1
Marinedrugs 21 00043 i006
Directly combine with Aβ to reduce Aβ in brain deposition[43]
Table 2. Main marine natural products with pharmacological activity to potentially treat AD.
Table 2. Main marine natural products with pharmacological activity to potentially treat AD.
FamilyCompoundOriginPharmacological ActivityRef.
PolysaccharideFucoidanBrown seaweedsBlock caspase-9 and caspase-3 enzymes.[57]
ChitosanCrustaceansInhibition of the enzyme acetylcholinesterase.[58]
MannanCodium fragileInhibition of the enzyme β-secretase (Bace1)[59]
PolyphenolDieckolBrown seaweeds
(Ecklonia cava, Ecklonia stolonifera)
Inhibition of the enzymes acetylcholinesterase and Butyrylcholinesterase[60]
SterolFucosterolbrown alarge
(Panida australis)
Anti-inflammatory and anti-BACE1[61,62]
CarotenoidFucoxanthinbrown algae
(Sargassum siliquastrum)
Anti-inflammatory and antioxidant[63]
Astaxanthingreen algae
(Haematococcus pluvialis)
Decrease the production of NF-κB transcription factors and inflammatory cytokines[64]
LycopeneRed seaweedsAnti-inflammatory and antioxidant[65]
DiterpenoidGracilinsMarine sponges
(Spongionella gracilis)
Inhibition of the enzyme β-secretase or BACE-1.Anti-inflammatory and antioxidant properties
Reduction in hyperphosphorylation of tau protein
[66]
AlkaloidManzamineMarine sponges (Acanthostrongylophora)GSK-3 inhibition and reduction in hyperphosphorylation of tau protein[67]
IndoleStreptomyces sp. Inhibition Aβ plaque production by activating the nuclear factor Nrf2[68]
Amino sulfonateHomotaurineRed seaweeds Aβ lowering and prevention of the formation of a toxic soluble amyloid oligomer[69,70]
Marine ToxinSpirolidesAlexandrium ostenfeldiiAcetylcholinesterase inhibition, and restraint the formation of amyloid-β[71]
MacrolideCaniferolide APhylum ActinobacteriaAnti-inflammatory and antioxidant action
Blockade of the BACE-1 enzyme.
[72]
Table 3. ADMET profiling of nine marine natural products.
Table 3. ADMET profiling of nine marine natural products.
Parameters 1FucoidanMannanFucosterolAstaxanthinLycopeneManzamine AIndoleHomotaurin13-desMeC
MW (130~725)242.05666.22412.37596.39536.44548.35117.06139.03691.44
LogS (−4~0.5)−0.0420.762−6.887−7.226−7.642−3.86−2.151−0.14−5.162
LogP (0~3)−2.003−4.8687.4478.04511.0725.4592.292−2.7455.652
Pgp-inh000.67910.9980.9990.0010.0010.998
Pgp-sub0.0060.8940.0010.0110.7580.0560.0120.0020.724
HIA0.92710.0040.0190.020.0340.0050.9250.027
F (30%)0.8510.2240.0010.1130.0030.4680.8350.534
Caco-2−5.53−6.294−4.624−5.196−5.708−5.131−4.259−5.902−4.813
BBB0.6780.450.8180.0010.0010.990.7370.9360.25
PPB18.86%5.85%98.64%101.00%99.44%96.16%86.32%10.54%97.60%
Fu72.18%54.83%1.78%2.56%5.42%3.27%16.50%87.62%1.54%
CYP1A2-inh0.00300.0580.0150.2660.280.9750.0050.005
CYP1A2-sub0.1710.0020.4360.1560.5210.7860.8050.2370.817
CL3.0240.2813.3040.719−0.2867.85111.1893.78218.565
T1/20.3020.5430.0160.0670.1370.0080.7940.5250.015
hERG0.0290.0250.0110.2350.8520.8630.0290.0350.679
Ames0.160.0720.0230.3690.3930.1540.3110.0560.02
ROA0.9430.0890.0220.1110.2180.8910.8440.3460.993
FDAMDD0.02700.6380.9740.9480.9320.2190.0140.942
BCF0.4440.1053.3171.4652.1871.0810.9340.2121.755
1 MW: Molecular weight. LogS: The logarithm of aqueous solubility value. LogP: The logarithm of the n-octanol/water distribution coefficient. Pgp-inh: The inhibitor of P-glycoprotein. Pgp-sub: The substrates of P-glycoprotein. HIA: Human intestinal absorption. F (30%): The human oral bioavailability 30%. Caco-2: The permeability of human colon adenocarcinoma cell lines (Caco-2). BBB: the penetration of blood–brain barrier (BBB). PPB: Plasma protein binding. Fu: The fraction unbound in plasms. CL: The clearance of a drug. T1/2: The half-life of a drug. hERG: The human ether-a-go-go related gene. Ames: The Ames test for mutagenicity. ROA: The toxicity of rat oral acute. FDAMDD: The maximum recommended daily dose. BCF: The bioconcentration factor. The data acquired from ADMETLab 2.0 database.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hu, D.; Jin, Y.; Hou, X.; Zhu, Y.; Chen, D.; Tai, J.; Chen, Q.; Shi, C.; Ye, J.; Wu, M.; et al. Application of Marine Natural Products against Alzheimer’s Disease: Past, Present and Future. Mar. Drugs 2023, 21, 43. https://doi.org/10.3390/md21010043

AMA Style

Hu D, Jin Y, Hou X, Zhu Y, Chen D, Tai J, Chen Q, Shi C, Ye J, Wu M, et al. Application of Marine Natural Products against Alzheimer’s Disease: Past, Present and Future. Marine Drugs. 2023; 21(1):43. https://doi.org/10.3390/md21010043

Chicago/Turabian Style

Hu, Di, Yating Jin, Xiangqi Hou, Yinlong Zhu, Danting Chen, Jingjing Tai, Qianqian Chen, Cui Shi, Jing Ye, Mengxu Wu, and et al. 2023. "Application of Marine Natural Products against Alzheimer’s Disease: Past, Present and Future" Marine Drugs 21, no. 1: 43. https://doi.org/10.3390/md21010043

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