Hostname: page-component-586b7cd67f-dsjbd Total loading time: 0 Render date: 2024-11-23T20:35:20.584Z Has data issue: false hasContentIssue false

Mitochondrial protective potential of fucoxanthin in brain disorders

Published online by Cambridge University Press:  25 July 2024

Khondoker Adeba Ferdous
Affiliation:
Department of Human Nutrition and Hospitality Management, College of Human Environmental Sciences, The University of Alabama, Tuscaloosa, AL, USA
Joseph Jansen
Affiliation:
Department of Human Nutrition and Hospitality Management, College of Human Environmental Sciences, The University of Alabama, Tuscaloosa, AL, USA
Emma Amjad
Affiliation:
Department of Human Nutrition and Hospitality Management, College of Human Environmental Sciences, The University of Alabama, Tuscaloosa, AL, USA
Eliana Pray
Affiliation:
Department of Human Nutrition and Hospitality Management, College of Human Environmental Sciences, The University of Alabama, Tuscaloosa, AL, USA
Rebecca Bloch
Affiliation:
Department of Human Nutrition and Hospitality Management, College of Human Environmental Sciences, The University of Alabama, Tuscaloosa, AL, USA
Alex Benoit
Affiliation:
Department of Human Nutrition and Hospitality Management, College of Human Environmental Sciences, The University of Alabama, Tuscaloosa, AL, USA
Meredith Callahan
Affiliation:
Department of Human Nutrition and Hospitality Management, College of Human Environmental Sciences, The University of Alabama, Tuscaloosa, AL, USA
Han-A Park*
Affiliation:
Department of Human Nutrition and Hospitality Management, College of Human Environmental Sciences, The University of Alabama, Tuscaloosa, AL, USA
*
*Corresponding author: Han-A Park, email: [email protected]

Abstract

Mitochondrial dysfunction is a common feature of brain disorders. Mitochondria play a central role in oxidative phosphorylation; thus changes in energy metabolism in the brain have been reported in conditions such as Alzheimer’s disease, Parkinson’s disease, and stroke. In addition, mitochondria regulate cellular responses associated with neuronal damage such as the production of reactive oxygen species (ROS), opening of the mitochondrial permeability transition pore (mPTP), and apoptosis. Therefore, interventions that aim to protect mitochondria may be effective against brain disorders. Fucoxanthin is a marine carotenoid that has recently gained recognition for its neuroprotective properties. However, the cellular mechanisms of fucoxanthin in brain disorders, particularly its role in mitochondrial function, have not been thoroughly discussed. This review summarises the current literature on the effects of fucoxanthin on oxidative stress, neuroinflammation, and apoptosis using in vitro and in vivo models of brain disorders. We further present the potential mechanisms by which fucoxanthin protects mitochondria, with the objective of developing dietary interventions for a spectrum of brain disorders. Although the studies reviewed are predominantly preclinical studies, they provide important insights into understanding the cellular and molecular functions of fucoxanthin in the brain. Future studies investigating the mechanisms of action and the molecular targets of fucoxanthin are warranted to develop translational approaches to brain disorders.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - ND
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives licence (http://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided that no alterations are made and the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use and/or adaptation of the article.
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of The Nutrition Society

Introduction

Brain disorders include a wide range of conditions such as neurodegenerative diseases, stroke, mental illness, epilepsy, traumatic brain injury (TBI), and cancer, and changes in cognition, movement, sense, and personality are commonly associated with these disorders. Data extracted from the Global Burden of Disease 2019 show that stroke is the leading cause of death and disability worldwide and that neurodegenerative diseases such as Alzheimer’s disease (AD) and Parkinson’s disease (PD) have increased considerably over the last 20 years.(Reference Ding, Wu and Chen1) Neurodegenerative diseases are generally correlated with aging. However, a recent study has reported an alarming increase in anxiety and depressive disorders among children and adolescents.(Reference Piao, Huang and Han2) Furthermore, researchers predict an increase in the incidence rate of ischaemic stroke in all age groups over the next decade.(Reference Pu, Wang and Zhang3)

Patients with brain disorders often live with symptoms or disabilities for an extended period. In patients with neurodegenerative diseases, structural or functional loss of brain cells can begin years or decades before clinical manifestations are exhibited,(Reference Sheinerman and Umansky4,Reference Katsuno, Sahashi and Iguchi5) and this process continues until death. Therefore, long-term interventions that are applicable to a wide range of the population may aid in the prevention or alleviation of the severity of brain disorders.(Reference Morris, Tangney and Wang6) Research suggests that diet is an important factor in the maintenance of brain function.(Reference Caracciolo, Xu and Collins7,Reference Marchand and Jensen8) Numerous studies have demonstrated the benefits of diets rich in vegetables, fruits, whole grains, and seafood for brain health. However, these dietary pattern studies generally do not include seaweed data.

The major objectives of this review are to introduce the neuroprotective properties of fucoxanthin, a bioactive compound found abundantly in brown seaweed, and to suggest potential cellular mechanisms based on recently published articles. In particular, we highlight the role of fucoxanthin in mitochondrial protection since mitochondrial dysfunction is a prominent and recurrent feature of various brain diseases, including AD, PD, and stroke. Here, we synthesise and analyse the existing scientific literature to determine the potential role of fucoxanthin in mitochondrial protection in the context of brain disorders.

Mitochondrial dysfunction in brain disorders

The brain requires a substantial amount of energy to maintain electrophysiological activities; therefore, the generation of ATP through oxidative phosphorylation by mitochondria is critical for brain health. Since the brain lacks sufficient long-term energy storage capacity, neurones are vulnerable to impaired energy metabolism,(Reference Wang, Zhao and Ma9) which commonly occurs in brain disorders that are associated with mitochondrial dysfunction. For example, the lack of blood flow to the brain disrupts the function of the electron transport chain (ETC), leading to energy failure in cases of cerebral ischaemia.(Reference Qin, Yang and Chu10) Furthermore, impaired energy metabolism has been reported in neurodegeneration models. The aggregation of α-synuclein inhibits the function of complex I in the ETC, which causes ATP depletion.(Reference Reeve, Ludtmann and Angelova11) Impaired ETC functioning has been reported in the brains of AD patients.(Reference Maurer, Zierz and Möller12)

ATP depletion impairs the ATP-dependent Na+/K+ and Ca2+ pumps, which leads to ionic imbalance in the cytosol, where increased Ca2+ ions activate cellular responses that cause neuronal damage. Ca2+ triggers the release of neurotransmitters into the extracellular space, including glutamate, which then activates N-methyl-D-aspartate receptors causing a large influx of Ca2+ and excitotoxicity. In addition, Ca2+ activates the opening of the mitochondrial permeability transition pore (mPTP),(Reference Robinson, Lee and DaSilva13,Reference Abramov and Duchen14) which is a large non-selective channel that triggers cell death. Once opened, mitochondrial membrane depolarization, ATP depletion, and mitochondrial swelling occur, ultimately leading to cell death. Research has shown that excitotoxicity triggers the dissociation of the F1 subcomplex from the Fo component of F1Fo ATP synthase, which results in the opening of a leak channel.(Reference Bonora, Morganti and Morciano15,Reference Mnatsakanyan, Park and Wu16) A rodent model of focal cerebral ischaemia showed a decreased mitochondrial membrane potential in isolated mitochondria when compared with the control model.(Reference Li, Ma and Yu17) An ischaemia-reperfusion model revealed that the loss of mitochondrial membrane potential was prevented by treatment with cyclosporine A, which inhibits mPTP.(Reference Liu, Wang and Zhang18) The neuroprotective properties of cyclosporine A were also observed in rodent TBI models.(Reference Sullivan, Thompson and Scheff19) In addition, amyloid-β (Aβ) peptide can trigger mPTP opening by causing the translocation of cyclophilin D.(Reference Du and Yan20) A deficiency in cyclophilin D could maintain mitochondrial membrane potential and improve cognitive function in AD mice, thereby indicating the involvement of mPTP in AD pathology.(Reference Du, Guo and Fang21) In a PD model, α-synuclein aggregates caused the loss of mitochondrial membrane potential and induced the activity of a leak channel similar to mPTP.(Reference Ludtmann, Angelova and Horrocks22)

Mitochondria are critical for the functioning of apoptotic pathways. Oligomerization of pro-apoptotic Bcl-2 family proteins such as Bax and Bak in the mitochondrial membrane promotes the release of cytochrome c from the mitochondria. This cytochrome c then binds to apoptotic protease-activating factor-1 and activates caspases,(Reference Kim, Du and Fang23) which causes the proteolysis of structural and functional proteins, leading to cell death. This apoptotic pathway can be inhibited by anti-apoptotic Bcl-2 family members, such as Bcl-2 and Bcl-xL, which sequester pro-apoptotic proteins. Maintaining a balance between pro- and anti-apoptotic Bcl-2 family proteins is critical for cell survival, and alterations in these protein levels are evident in brain tissues of humans and animals with brain disorders. Postmortem studies have identified increased levels of pro-apoptotic Bax proteins in the brain tissues of PD patients.(Reference Tatton24) Similarly, mouse PD models have shown increased pro-apoptotic Bax protein and mRNA levels and decreased anti-apoptotic Bcl-2 protein levels in dopaminergic neurones.(Reference Vila, Jackson-Lewis and Vukosavic25) In addition, genetically modified Bax-deficient PD mice were revealed to be resistant to the loss of dopaminergic neurones.(Reference Vila, Jackson-Lewis and Vukosavic25) The Aβ peptide, which is a hallmark of AD, can decrease Bcl-2 and increase Bax, thereby potentially contributing to neuronal apoptosis and neurodegeneration.(Reference Paradis, Douillard and Koutroumanis26) Increased levels of Bax and decreased levels of Bcl-2 and Bcl-xL have been observed in a rodent model of cerebral ischaemia,(Reference Zhang and Wang27) and treatment with pro-apoptotic protein inhibitors reversed excitotoxicity-associated neuronal death.(Reference Park, Licznerski and Mnatsakanyan28)

Mitochondria are major contributors to cellular ROS, which are by-products of oxidative phosphorylation.(Reference Lin and Beal29) ETC impairment can cause excessive ROS production, which overwhelms the antioxidant defence system and promotes oxidative stress. Increases in oxidative stress markers that indicate lipid peroxidation and DNA oxidation have been observed in the brain during neurodegeneration(Reference Butterfield and Halliwell30,Reference Su, Wang and Nunomura31) and the depletion of antioxidant enzymes such as glutathione peroxidase and superoxide dismutase (SOD) has been identified in neurodegenerative diseases. In AD, Aβ plaques accumulate in the brain and disrupt the activities of ETC complexes, which causes an overproduction of ROS.(Reference Maurer, Zierz and Möller12) Moreover, oligomeric or fibrillar α-synuclein inhibits the function of the mitochondrial respiratory chain, resulting in increased ROS production in PD cases.(Reference Reeve, Ludtmann and Angelova11) Oxidative stress can activate the NLRP3 inflammasome, which is a multiprotein complex that triggers the processing and secretion of pro-inflammatory cytokines.(Reference Sorbara and Girardin32) Elevated levels of ROS and ROS-mediated cellular damage in neurodegenerative diseases trigger inflammatory responses that cause immune cell recruitment and additional ROS production. Extracellular accumulation of Aβ plaques can activate microglia, which release pro-inflammatory cytokines.(Reference Wang, Tan and Yu33) In cerebral ischaemia, both hypoxia and reperfusion trigger inflammatory responses that activate macrophages and release pro-inflammatory cytokines, including TNF alpha (TNF-α), IL-1, IL-6, and IL-8.(Reference Qin, Yang and Chu10) These inflammatory cytokines induce the production of adhesion molecules on endothelial cells that promote neutrophil adhesion; these activated neutrophils then generate additional ROS.(Reference Kawabori and Yenari34)

Mitochondrial DNA has been implicated in the pathogenesis of neurodegenerative disorders because of its predisposition to oxidative damage due to its proximity to the ETC and the lack of a proper DNA repair mechanism.(Reference Chen, Turnbull and Reeve35) Furthermore, mutations in genes associated with mitochondrial function have been linked to neurodegeneration, such as the PARK genes, which are critical for mitochondrial function and have been implicated in the pathogenesis of PD. Mutations in the PINK1 and Parkin genes have been associated with decreased ETC enzyme activity and ATP depletion.(Reference Chen, Turnbull and Reeve35) These mutations disrupt the balance of mitochondrial fission, fusion, and mitophagy, leading to an increased number of fragmented mitochondria. Mutations in the DJ1 and LRRK2 genes are associated with decreased mitochondrial membrane potential, increased oxidative stress, decreased complex I function, and increased mitochondrial fragmentation.(Reference Lev, Roncevich and Ickowicz36,Reference Martin, Kim and Dawson37)

Neuroprotective effects of fucoxanthin

Fucoxanthin (C42H58O6) is a carotenoid marine xanthophyll that is found in brown seaweed such as Undaria pinnatifida and Laminaria japonica.(Reference Wang, Wang and Zhang38,Reference Fung, Hamid and Lu39) It has a strong antioxidant capacity because of the presence of multiple double bonds, including a conjugated carbonyl group. Fucoxanthin has a radical scavenging effect when treated with 1,1-diphenyl-2-picrylhydrazyl and 2,2′-Azinobis-3-ethylbenzo thiazoline-6-sulfonate(Reference Nomura, Kikuchi and Kubodera40Reference Sachindra, Sato and Maeda43) and exhibits a stronger hydroxyl radical quenching activity than other antioxidants.(Reference Sachindra, Sato and Maeda43) In addition to its direct scavenging effect, fucoxanthin regulates the gene expression of antioxidant enzymes such as SOD and catalase to help support cellular redox homeostasis.(Reference Xiang, Liu and Lin44)

Although epidemiological data demonstrating the direct effect of fucoxanthin on brain disorders are limited, a high intake of seaweed has been reported to lower the risk of PD(Reference Okubo, Miyake and Sasaki45) and stroke-mediated mortality in humans.(Reference Kishida, Yamagishi and Muraki46) Pharmacokinetic studies using human plasma detected fucoxanthinol, a metabolite of fucoxanthin, after a single oral administration of kombu (Laminaria japonica) extract containing fucoxanthin (31 mg)(Reference Hashimoto, Ozaki and Mizuno47) or a 1-week dose of dried wakame (Undaria pinnatifida) containing 6.1 mg of fucoxanthin.(Reference Asai, Yonekura and Nagao48) Quantitative analyses of fucoxanthin in human tissues other than blood have not been conducted; however, Hashimoto et al. observed its distribution and that of its metabolites such as fucoxanthinol and amarouciaxanthin A in tissues after oral administration in a mouse model.(Reference Hashimoto, Ozaki and Taminato49) Although the brain was not analysed in that study, fucoxanthin was present in other lipid-rich organs such as the adipose and liver. While the neuroprotective potential of fucoxanthin has been documented in in vivo animal models, its ability to cross the blood-brain barrier (BBB) remains unconfirmed. However, since other carotenoids from the xanthophyll family, such as astaxanthin, lutein, and zeaxanthin, can permeate the BBB, fucoxanthin has the potential for uptake into brain tissue.(Reference Manabe, Komatsu and Seki50,Reference Vishwanathan, Neuringer and Snodderly51)

Alzheimer’s disease

Alzheimer’s disease is a progressive neurodegenerative disorder that is characterised by cognitive impairment, memory decline, and behavioral and personality changes. It is the most common cause of dementia in older adults and affects one in nine US residents aged 65 and older.(52) The complex neuropathology of AD involves the extracellular aggregation of Aβ into plaques and the intracellular accumulation of hyperphosphorylated tau protein into neurofibrillary tangles. These processes accompany and affect significant mitochondrial dysfunction in AD models. Aβ can impair the functioning of ETC, cause the loss of mitochondrial membrane potential, increase ROS production, and disrupt Ca2+ homeostasis.(Reference ME53Reference Kuchibhotla, Goldman and Lattarulo56) In addition to mitochondrial dysfunction, Aβ induces neuroinflammation and impairs neurotransmission.

Fucoxanthin has been suggested to protect mitochondria against AD-associated pathologies.(Reference Xiang, Liu and Lin44,Reference Lin, Yu and Zhao57,Reference Lee, Youn and Yoon58) Treatment with fucoxanthin prevents the loss of cell viability against Aβ-induced cytotoxicity.(Reference Lin, Yu and Zhao57,Reference Lee, Youn and Yoon58) In particular, fucoxanthin-treated cells were resistant to apoptotic death, thereby indicating mitochondrial protection,(Reference Lin, Yu and Zhao57,Reference Lee, Youn and Yoon58) and fucoxanthin-modified Aβ1-42 oligomers have been reported to be less toxic than Aβ1-42 oligomers to SH-SY5Y cells.(Reference Xiang, Liu and Lin44) Furthermore, fucoxanthin may directly prevent the formation of Aβ plaques and neurofibrillary tangles. Xiang et al. showed the binding of fucoxanthin to Aβ1-42 peptides, where it inhibited the formation of Aβ fibrils and oligomers(Reference Xiang, Liu and Lin44) and prevented Aβ-mediated mitochondrial dysfunction.(Reference Itoh, Wakabayashi and Katoh59Reference Dickson62) Similarly, Jung et al. reported the binding of fucoxanthin and inhibition of β-site amyloid precursor protein cleaving enzyme 1, which cleaves the amyloid precursor protein to produce Aβ.(Reference Jung, Ali and Choi63) Fucoxanthin may inhibit the production and aggregation of Aβ through interaction with two hydroxyl groups.(Reference Jung, Ali and Choi63) In addition, co-incubation of Aβ monomers with fucoxanthin resulted in a dose-dependent decrease in Aβ oligomer formation through hydrophobic interactions. Lee et al. reported that fucoxanthin reversed the loss of mitochondrial membrane potential in Aβ-treated PC12 cells.(Reference Lee, Youn and Yoon58) In that study, 5 μM fucoxanthin was as effective as 50 μM resveratrol, an antioxidant with neuroprotective function, indicating a strong mitochondrial protection capacity. Fucoxanthin inhibits Aβ-mediated upregulation of Bax, thereby helping to maintain the integrity of the mitochondrial membrane.(Reference Lee, Youn and Yoon58) Those authors also showed that treatment with fucoxanthin prevented the increase of intracellular Ca2+, as measured by fluo3-AM. Although mPTP was not primarily discussed in that study, the data suggests that fucoxanthin has a role in inhibiting the opening of the mPTP.

Fucoxanthin protects cells from oxidative damage associated with AD.(Reference Lin, Yu and Zhao57,Reference Lee, Youn and Yoon58) Treatment with fucoxanthin was shown to increase nuclear Nrf2 expression, whereas co-treatment with a PI3K inhibitor attenuated this increase, suggesting that fucoxanthin mediated this increased expression through the Akt/GS3K signalling pathway.(Reference Lee, Youn and Yoon58) Similarly, treatment with fucoxanthin increased the activation of the pro-survival PI3K/Akt pathway in Aβ SH-SY5Y cells.(Reference Lin, Yu and Zhao57) Nrf2 is an important regulator of cellular antioxidants that is normally sequestered in the cytosol by Keap1.(Reference Itoh, Wakabayashi and Katoh59) Oxidative stress induces the translocation of Nrf2 to the nucleus, where pFyn eventually stimulates its export. Fucoxanthin may increase nuclear Nrf2 by preventing the phosphorylation of Fyn and Nrf2 by GS3K, thereby decreasing the degradation and export of Nrf2 from the nucleus.(Reference Lee, Youn and Yoon58) Similarly, fucoxanthin regulated redox homeostasis in in vivo models of AD and was shown to attenuate the decrease in SOD, catalase, and glutathione in AD mice.(Reference Xiang, Liu and Lin44) The BV2 cells treated with the poly lactic-co-glycolic acid-block-polyethylene glycol (PLGA-PEG)-fucoxanthin nanoparticle prevented Aβ-induced NF-κB, TNF-α, and IL-1β induction compared with the BV2 cells treated with only Aβ oligomers, indicating that fucoxanthin prevented Aβ-induced neuroinflammation.(Reference Yang, Jin and Wu60) Furthermore, fucoxanthin has anticholinesterase and antibutyrylcholinesterasic activity,(Reference Kawee-ai, Kuntiya and Kim61) and since acetylcholine is considerably depleted in the AD pathology, fucoxanthin may help maintain acetylcholine levels in the brain.(Reference Kawee-ai, Kuntiya and Kim61)

Parkinson’s disease

Parkinson’s disease is a progressive neurodegenerative disorder caused by the loss of dopaminergic neurones in the substantia nigra that leads to a deficiency of dopamine, which is a neurotransmitter that controls movement, pleasure, and motivation.(Reference Dickson62) The common symptoms of PD include tremors, rigidity, akinesia, and postural instability,(Reference Dickson62) while non-motor symptoms such as depression, anxiety, and cognitive impairments can also occur.(Reference Dickson62) In 2019, over 8.5 million individuals worldwide were estimated to be living with PD.(64) A 2022 study stated that close to 90,000 people are diagnosed with PD every year in the United States, representing a 50% increase from the previously estimated number of 60,000 annually.(65) Currently, patients with PD are treated with levodopa (L-DA) to alleviate symptoms, although this treatment may cause neurotoxicity in the long term.(Reference Tambasco, Romoli and Calabresi66)

The degradation of dopaminergic neurones is associated with the accumulation of misfolded α-synuclein proteins. Oligomeric and fibrillar α-synucleins inhibit the function of ETC and impair ATP production. Postmortem studies have shown decreases in complexes I and II in the PD brain.(Reference Grünewald, Rygiel and Hepplewhite67) Furthermore, the α-synuclein aggregate interacts with F1Fo ATP synthase in the respiratory chain, which causes the opening of the mPTP,(Reference Ludtmann, Angelova and Horrocks22) and disruption of the respiratory chain leads to the overproduction of ROS that damages the mitochondria. Increased oxidative stress also impairs the function of the ubiquitin-proteasome system, thereby inhibiting the clearance of these misfolded proteins and leading to their accumulation in neurones.(Reference Moon and Paek68) Degradation of dysfunctional mitochondria is important for maintaining a healthy pool of neuronal mitochondria. Mutations in the PINK1 and Parkin genes in the PD brain impair the mitophagy process resulting in the accumulation of dysfunctional mitochondria.(Reference Pickrell and Youle69) Similarly, DJ1 and LRRK2 gene mutations are associated with loss of mitochondrial membrane potential and morphology.(Reference Lev, Roncevich and Ickowicz36,Reference Martin, Kim and Dawson37) DJ1 binds to the β-subunit of the F1Fo ATP synthase enhancing ATP production. In contrast, a DJ1 mutant fails to close the mitochondrial inner membrane leak, thereby altering energy metabolism.(Reference Chen, Park and Mnatsakanyan70) The aggregation of α-synuclein can also affect mitochondrial fragmentation(Reference Nakamura, Nemani and Azarbal71) and movement, which alters the mitochondrial distribution in axonal regions that have high energy demands.(Reference Thorne and Tumbarello72)

In vitro and in vivo studies have suggested the mitochondrial protective effects of fucoxanthin in PD pathology. Treatment with fucoxanthin prevented the loss of mitochondrial membrane potential in PC12 cells challenged with 6-hydroxydopamine (6-OHDA) or a combination of 6-OHDA and L-DA.(Reference Wu, Han and Liu73,Reference Liu, Lu and Tang74) Although the levels of ATP production and oxygen consumption were not directly measured in those models, maintaining mitochondrial membrane potential is critically important to power F1Fo ATP synthase. Thus, fucoxanthin may help maintain mitochondrial energy metabolism. In addition, annexin V and propidium iodide co-staining showed that fucoxanthin prevented PD-associated apoptotic cell death,(Reference Wu, Han and Liu73,Reference Liu, Lu and Tang74) suggesting that fucoxanthin is involved in the protection of mitochondrial membrane integrity. The direct role of fucoxanthin on mitochondrial quality control in the PD brain is unknown. Lian et al. showed that treatment with fucoxanthin increased the ratio of LC3-II to LC3-I, the protein level of Parkin, and the number of autophagosomes and mitophagosomes in retinal ganglion cells challenged with excitotoxicity,(Reference Lian, Hu and Zhang75) suggesting that fucoxanthin has a role in regulating mitophagy. Those authors further showed that fucoxanthin prevented the loss of mitochondrial membrane potential during excitotoxicity and helped protect from apoptotic death by lowering Bax and increasing Bcl-2.

The loss of mitochondrial membrane integrity and inefficient operation of ETC increases ROS production. PC12 cells challenged with 6-OHDA or a combination of 6-OHDA and L-DA showed increased DCF signals indicating increased intracellular ROS,(Reference Wu, Han and Liu73,Reference Liu, Lu and Tang74) whereas fucoxanthin-treated PC12 cells were resistant to ROS production in a dose-dependent manner. In addition to its radical scavenging properties,(Reference Nomura, Kikuchi and Kubodera40Reference Yan, Chuda and Suzuki42) fucoxanthin increased antioxidant enzyme expression in PD models.(Reference Wu, Han and Liu73,Reference Liu, Lu and Tang74) Fucoxanthin binds to a hydrophobic site on Keap1 where it decreases the affinity of Keap1 to Nrf2-binding in a dose-dependent manner.(Reference Wu, Han and Liu73) Therefore, fucoxanthin increased the nuclear expression of Nrf2 and the downstream genes that encode antioxidant enzymes such as haem oxygenase-1, the glutamate-cysteine ligase modifier subunit, and the glutamate-cysteine ligase catalytic subunit in 6-OHDA-treated PC12 cells.(Reference Wu, Han and Liu73) In addition, PD mice that underwent intragastric administration of fucoxanthin (50, 100, or 200 mg/kg/day) for 28 d improved pole climbing, swimming, and suspension experiment scores, indicating improved motor function.(Reference Liu, Lu and Tang74) Similarly, zebrafish larvae treated with different concentrations of fucoxanthin for 4 d showed improved swimming abilities after exposure to 6-OHDA.(Reference Wu, Han and Liu73)

Cerebral ischaemia

Cerebral ischaemia is a condition that commonly occurs during cardiovascular events such as stroke or cardiac arrest and involves reduced blood flow to the brain. This condition impairs mitochondrial energy metabolism through the deprivation of oxygen and essential nutrients to the brain.(Reference Sims and Mitochondria76,Reference Yang, Mukda and Chen77) Dysfunctional mitochondria cause the production of excessive ROS, which damage various cellular components and trigger apoptotic pathways, including the oligomerization of pro-apoptotic proteins in the mitochondrial membrane and the release of cytochrome c, leading to programmed cell death.(Reference Broughton, Reutens and Sobey78) In addition, ischaemia-mediated energy depletion causes the failure of the ATP-dependent ionic pump, thereby altering intracellular ionic homeostasis. Furthermore, excessive cellular Ca2+ concentrations cause the opening of mPTP,(Reference Li, Park and Jonas79,Reference Baumgartner, Gerasimenko and Thorne80) which exacerbates the energy crisis and leads to neuronal death.

Ikeda et al. showed that treatment with fucoxanthin isolated from wakame (Undaria pinnatifida) promoted the release of lactate dehydrogenase in hypoxia-exposed primary cortical neurones,(Reference Ikeda, Kitamura and Machida81) which suggests that fucoxanthin reduces cytotoxicity during oxygen depletion. Those authors performed an in vivo study using stroke-prone spontaneously hypertensive rats, which were characterised by severe spontaneous hypertension and the development of cerebrovascular diseases. Supplementation with 5% wakame powder delayed the development of stroke and increased the lifespan of the rats. Hu et al. further investigated the cellular mechanisms of fucoxanthin-mediated neuroprotection(Reference Hu, Chen and Tian82) by intragastrically administering 30, 60, and 90 mg/kg of fucoxanthin to Wistar rats 1 h before middle cerebral artery occlusion (MCAO). The results showed that rats treated with fucoxanthin exhibited a dose-dependent reduction of MCAO-induced brain injury. Treatment with fucoxanthin increased the ratio of Bcl-2/Bax and decreased the cleaved caspase 3 protein level, indicating inhibition of mitochondria-mediated apoptosis during cerebral ischaemia. Consistent with in vivo data, rat cortical neurones treated with fucoxanthin showed anti-apoptotic properties in response to oxygen-glucose deprivation and reoxygenation challenges.(Reference Hu, Chen and Tian82) Furthermore, the study suggested that fucoxanthin increased antioxidant proteins such as SOD(Reference Hu, Chen and Tian82) via the activation of Nrf2,(Reference Wang, Zhang and Lu83) thereby protecting neurones from oxidative stress. Wang et al. used PLGA-PEG nanoparticles to increase fucoxanthin bioavailability in the brains of MCAO-induced rats.(Reference Wang, Zhang and Lu83) PLGA-PEG encapsulation improves fucoxanthin stability in the body and allows for its extended release and enhanced penetration into the central nervous system. The results of that study showed that the intravenous administration of PLGA-PEG fucoxanthin nanoparticles (20 and 40 mg/kg) half an hour before MCAO reduced the behavioural deficits associated with cerebral ischaemia in the rats. In addition, the infarct volumes and brain oedema extents were decreased in rats receiving the nanoparticle treatment.(Reference Wang, Zhang and Lu83) Furthermore, PLGA-PEG fucoxanthin nanoparticles prevented the loss of glutathione peroxidase, SOD, and catalase activity in the ischaemic brain indicating the roles of these compounds in regulating antioxidant defence. Fucoxanthin nanoparticles exhibit anti-inflammatory properties through the inactivation of the NF-κB pathway.

Depression and anxiety

Depression is characterised by a pervasive lack of interest in daily activities that can lead to a profound sense of hopelessness or self-harm, while anxiety involves excessive worry and enduring fear. Depression and anxiety frequently co-occur, possibly due to overlapping cellular mechanisms, with 41.6% of individuals diagnosed with a depressive episode presenting with anxiety within 12 months of the diagnosis,(Reference Kalin84) demonstrating a substantial comorbidity between the two conditions. Although the neurological pathways that affect these disorders are not fully understood, a growing body of studies suggests an association with mitochondrial dysfunction.(Reference Fox and Lobo85,Reference Hollis, Pope and Gorman-Sandler86) Cellular energy metabolism is dysregulated in patients with depression.(Reference Li, Su and Wang87) Similarly, a proteomic analysis of mice challenged with chronic corticosterone, a stress hormone associated with depression and anxiety, showed that oxidative phosphorylation-related protein expression was decreased in these mice.(Reference Gong, Yan and Lei88) Treatment with ATP reversed the impaired synaptic transmission and excitability in neurones in depression mouse models.(Reference Lin, Huang and Luo89) Moreover, repeated unpredictable stress downregulated anti-apoptotic genes such as Bcl-2 and Bcl-xL in rat brains,(Reference Kosten, Galloway and Duman90) whereas approaches that improved the anti-apoptotic/pro-apoptotic Bcl-2 protein ratio alleviated the depression-associated behaviours.(Reference Wang, Xie and Zhang91)

Although the risk factors for behavioural disorders are complex, studies have demonstrated that the consumption of a healthy diet lowers the risk of depression.(Reference Kris-Etherton, Petersen and Hibbeln92,Reference Molendijk, Molero and Sánchez-Pedreño93) A prospective cohort study with Japanese adults, which was adjusted for biological, socio-economic, and dietary factors, reported that high seaweed intake was negatively associated with depressive symptoms.(Reference Guo, Huang and Cui94) In animal models, treatment with extract from the brown seaweed Sargassum horneri (500 mg/kg) for 3 weeks prevented the loss of neurotransmitters such as serotonin, dopamine, and norepinephrine in the mouse brain, as well as improvements in depressive-live behaviours caused by an intraperitoneal injection of corticosterone.(Reference Park, Kim and Kim95) The authors performed a quantitative analysis and verified the presence of fucoxanthin in the extract. That study further revealed that the underlying mechanism involved the activation of the ERK-CREB-BDNF pathways by the brown seaweed extract. Although the role of Sargassum horneri in mitochondrial protection was not intensively featured in that study, the BDNF has been previously shown to increase anti-apoptotic Bcl-xL(Reference Chao, Ma and Lee96) where the depletion of Bcl-xL impairs the maturity of BDNF.(Reference Park, Crowe-White and Ciesla97) Thus, treatment with fucoxanthin may promote the development and growth of neurones by preventing the mitochondrial dysfunction associated with depression. In addition, the intragastric administration of fucoxanthin (0, 50, 100, and 200 mg/kg) improved Lipopolysaccharide (LPS)-induced anxiety behaviours in mice.(Reference Jiang, Wang and Lin98) Treatment with fucoxanthin regulates the AMPK-NF-κB pathways and prevents the accumulation of pro-inflammatory cytokines, such as IL-1β, IL-6, and TNF-α, and iNOS and COX-2 in the hippocampus, cortex, and hypothalamus.

Brain aging

The cumulative impact of oxidative damage on neurones contributes to the process of brain aging.(Reference Harman99,Reference Paradies, Petrosillo and Paradies100) This oxidative damage results from the generation of free radicals through environmental exposure, inflammation, immune responses, and infection.(Reference Harman99,Reference Paradies, Petrosillo and Paradies100) Postmortem analyses of aged human brain tissues, particularly in the hippocampus and cortical regions, revealed elevated ROS levels and a decrease in antioxidant enzymes such as SOD and catalase.(Reference Venkateshappa, Harish and Mahadevan101) The accumulation of dysfunctional mitochondria in aged brains is associated with brain atrophy and neuronal apoptosis,(Reference Boveris and Navarro102) and the mitochondria isolated from the brains and other tissues of aged rodents demonstrated decreased ETC activity. Specifically, ETC complexes I and IV exhibit reduced activity in the aging brain, which can be attributed to increased oxidative stress.(Reference Boveris and Navarro102)

The diet patterns in Okinawa, a blue zone region with a high life expectancy, include nutrient-dense foods such as lean meat, fish, and vegetables, including seaweed.(Reference Willcox and Willcox103) Fucoxanthin is abundant in brown seaweed, which is a common part of the Okinawan diet. In the 1990s, Okinawan people had the lowest mortality rate from age-associated diseases in the Japanese and US populations.(Reference Willcox, Willcox and Todoriki104) Research investigating the direct role of fucoxanthin in brain aging is still limited; however, oxidative stress and inflammation are the main risk factors for age-associated cognitive decline,(Reference Baierle, Nascimento and Moro105,Reference Ownby106) and the antioxidant effects of fucoxanthin have been determined.(Reference Nomura, Kikuchi and Kubodera40Reference Yan, Chuda and Suzuki42) Chen et al. reported that the administration of 100–200 mg/kg of fucoxanthin for 7 d post-laparotomy surgery improved cognitive function in 12–14-month-old mice.(Reference Chen, Dong and Gong107) This treatment also decreased neuroinflammation and oxidative stress by activating the Akt and Nrf2 pathways, decreasing the levels of pro-inflammatory cytokines such as TNF-α and IL-1β, and increasing the antioxidant enzymes in the hippocampal tissue.(Reference Chen, Dong and Gong107)

Fucoxanthin increased the lifespan and resilience of Caenorhabditis elegans (C. elegans) and Drosophila melanogaster models to starvation and thermal and oxidative stress. Two carotenoids, fucoxanthin and β-carotene, were tested in that study, and lifespan extensions in C. elegans were only observed in the fucoxanthin-treated group.(Reference Lashmanova, Proshkina and Zhikrivetskaya108) Additionally, transcriptome analysis showed that fucoxanthin regulates the pathways involved in longevity, Wnt, and autophagy.(Reference Moskalev, Shaposhnikov and Zemskaya109) In particular, the Wnt signalling pathway controls mitochondrial function, including metabolism, biogenesis, and dynamics in cancer and non-transformed cells. Increased Wnt signalling activates mitophagy by increasing ROS production, which in turn decreases the number of damaged mitochondria and increases the quantity of working mitochondria through biogenesis. The Wnt signalling pathway is important for maintaining mitochondrial homeostasis; therefore, fucoxanthin-mediated upregulation of genes associated with the Wnt pathway could lead to mitochondrial protection. In addition, autophagy is an important quality control mechanism to remove damaged organelles such as dysfunctional mitochondria. Dysregulated autophagy promotes protein aggregation, which is associated with neurodegeneration such as that of PD. Furthermore, this study also showed that oxidative phosphorylation and apoptosis were one of the major targets of fucoxanthin, suggesting its role in regulating mitochondrial function.

Discussion

Mitochondrial protection is critically important to prevent the ATP depletion that occurs as an effect of brain disorders. In addition to cellular energy metabolism, targeting mitochondria can prevent pathological cell damage associated with ROS, apoptosis, and mPTP opening.(Reference Dagda110) Fucoxanthin demonstrates neuroprotective effects towards a range of brain disorders including AD, PD, cerebral ischaemia, depression, and anxiety (Table 1). Although limited studies have focused on the direct mechanisms of fucoxanthin-mediated mitochondrial function, fucoxanthin has been implicated in mitochondrial protection in several reports (Fig. 1). In particular, mitochondria-mediated apoptosis under fucoxanthin treatment has been tested in various brain disorder models. It was found that this treatment prevented the loss of the Bcl-2/Bax ratio and the cleavage of caspase-3 in ischaemia-induced brain infarct.(Reference Hu, Chen and Tian82) Similar results have been reported in cells challenged with Aβ and 6-OHDA, in which the anti-apoptotic Bcl-2 family protein was maintained while pro-apoptotic proteins were suppressed.(Reference Lee, Youn and Yoon58,Reference Liu, Lu and Tang74) Treatment with fucoxanthin has also been shown to prevent Aβ-induced intracellular Ca+2 increase in PC12 cells,(Reference Lee, Youn and Yoon58) suggesting the potential role of this compound in halting the opening of the mPTP. Despite promising data regarding the mitochondrial protective roles of fucoxanthin, the measurements that show the effects on cellular energy metabolism, such as those of ATP production and oxygen consumption, have not been performed in brain disorder models. However, cells treated with fucoxanthin have been shown to maintain rhodamine 123 and JC-1 fluorescent signals in neurodegenerative disease models, indicating the presence of the mitochondrial membrane potential.(Reference Lee, Youn and Yoon58,Reference Liu, Lu and Tang74) Since the mitochondrial membrane potential drives the operation of F1Fo ATP synthase, fucoxanthin may help alleviate the neuronal dysfunction that is associated with energy depletion in brain disorders.

Table 1. The effects of fucoxanthin on in vitro and in vivo brain disorder models

HPA, hypothalamic-pituitary-adrenal; ICR, Institute of Cancer Research; ROS, reactive oxygen species; SHRSP, Stroke-prone spontaneously hypertensive.

Fig. 1. Mechanism of mitochondrial protection by fucoxanthin in brain disorders.

Oxidative stress and inflammation, which commonly occur in brain disorders during pathological processes, can damage intracellular organelles including mitochondria; therefore, the antioxidant and anti-inflammatory effects of fucoxanthin (Fig. 1) have been studied with various models. In particular, treatment with fucoxanthin has been shown to activate Nrf-2 signalling and upregulate genes that encode antioxidant enzymes such as SOD, catalase, and haem oxygenase-1 in conditions with neurotoxic challenges induced by Aβ, 6-OHDA, and ischaemia.(Reference Lee, Youn and Yoon58,Reference Yang, Jin and Wu60,Reference Wu, Han and Liu73) Mitochondria are the major sources of ROS, and the fucoxanthin-mediated transcriptional regulation of antioxidants may protect mitochondria from oxidative stress. In addition, fucoxanthin has been shown to alleviate neuroinflammation in the hippocampus and cortex of Aβ-treated mice.(Reference Yang, Jin and Wu60) This study showed that fucoxanthin decreased the NF-κB activity and its downstream targets such as TNF-α and IL-1β. Similarly, treatment with fucoxanthin prevented NF-κB p65 expression in the hippocampus, cortex, and hypothalamus of LPS-injected depressive mice(Reference Jiang, Wang and Lin98) and the production of TNF-α in MCAO-induced rats.(Reference Wang, Zhang and Lu83)

In this review, we discuss the existing literature on the potential of fucoxanthin to protect mitochondria in brain disorders. However, we acknowledge limitations. First, the existing body of literature investigating the mechanism of fucoxanthin on mitochondrial function in the context of brain diseases is limited. This scarcity of dedicated studies hinders a comprehensive understanding of the specific effects of fucoxanthin on mitochondrial function in the brain. In addition, the majority of the discussed evidence was derived from preclinical studies. While the results from these experimental studies provide valuable insights into the cellular and molecular aspects, the findings must be adaptable to clinical applications in humans. Moreover, the bioavailability of fucoxanthin could pose a challenge since the absorption, distribution, metabolism, and excretion in the human body may influence its efficacy; therefore, these aspects should be explored.

Conclusion

Fucoxanthin has been shown to protect the brain against challenges associated with brain disorders. Here, we discuss the potential roles of fucoxanthin in cellular responses associated with mitochondrial dysfunction. The antioxidant and anti-apoptotic effects of fucoxanthin suggesting mitochondrial protection have been evaluated in various brain disorder models. However, future mechanistic studies focusing on the role of fucoxanthin in mitochondrial function, along with clinical studies on its efficacy in alleviating neuronal damage, can aid in the development of dietary recommendations to offset the burden of brain disorders.

Financial support

The authors did not receive any financial support from a funding agency, commercial sector, or non-profit organisation.

Competing interests

The authors declared none.

Authorship

KAF was involved in drafting the original manuscript, interpretation of findings, and editing of the manuscript. JJ was involved in drafting the original manuscript, interpretation of findings, and editing of the manuscript. EA was involved in drafting the original manuscript, interpretation of findings, and editing of the manuscript. EP was involved in drafting the original manuscript, interpretation of findings, and editing of the manuscript. RB was involved in drafting the original manuscript, interpretation of findings, and editing of the manuscript. AB was involved in drafting the original manuscript, interpretation of findings, and editing of the manuscript. MC was involved in drafting the original manuscript. HP was involved in drafting the original manuscript, interpretation of findings, and editing of the manuscript.

References

Ding, C, Wu, Y, Chen, X, et al. Global, regional, and national burden and attributable risk factors of neurological disorders: the Global Burden of Disease study 1990–2019. Front Public Health. 2022;10:952161.CrossRefGoogle ScholarPubMed
Piao, J, Huang, Y, Han, C, et al. Alarming changes in the global burden of mental disorders in children and adolescents from 1990 to 2019: a systematic analysis for the Global Burden of Disease study. Eur Child Adolesc Psychiatry. 2022;31(11):18271845.CrossRefGoogle ScholarPubMed
Pu, L, Wang, L, Zhang, R, et al. Projected global trends in ischemic stroke incidence, deaths and disability-adjusted life years from 2020 to 2030. Stroke. 2023;54(5):13301339.CrossRefGoogle ScholarPubMed
Sheinerman, KS, Umansky, SR. Early detection of neurodegenerative diseases: circulating brain-enriched microRNA. Cell Cycle. 2013;12(1):12.CrossRefGoogle ScholarPubMed
Katsuno, M, Sahashi, K, Iguchi, Y, et al. Preclinical progression of neurodegenerative diseases. Nagoya J Med Sci. 2018;80(3):289298.Google ScholarPubMed
Morris, MC, Tangney, CC, Wang, Y, et al. MIND diet slows cognitive decline with aging. Alzheimer’s dementia. 2015;11(9):10151022.CrossRefGoogle ScholarPubMed
Caracciolo, B, Xu, W, Collins, S, et al. Cognitive decline, dietary factors and gut-brain interactions. Mech Ageing Dev. 2014;136–137:5969.CrossRefGoogle ScholarPubMed
Marchand, NE, Jensen, MK. The role of dietary and lifestyle factors in maintaining cognitive health. Am J Lifestyle Med. 2018;12(4):268285.CrossRefGoogle ScholarPubMed
Wang, W, Zhao, F, Ma, X, et al. Mitochondria dysfunction in the pathogenesis of Alzheimer’s disease: recent advances. Mol Neurodegener 2020;15:122.CrossRefGoogle ScholarPubMed
Qin, C, Yang, S, Chu, Y-H, et al. Signaling pathways involved in ischemic stroke: molecular mechanisms and therapeutic interventions. Signal Transduction Targeted Therapy. 2022;7(1):129.Google ScholarPubMed
Reeve, A, Ludtmann, MH, Angelova, P, et al. Aggregated α-synuclein and complex I deficiency: exploration of their relationship in differentiated neurons. Cell death disease. 2015;6(7):e1820-e.CrossRefGoogle ScholarPubMed
Maurer, I, Zierz, S, Möller, HJ. A selective defect of cytochrome c oxidase is present in brain of Alzheimer disease patients. Neurobiol Aging 2000;21(3):455462.CrossRefGoogle ScholarPubMed
Robinson, MB, Lee, ML, DaSilva, S. Glutamate transporters and mitochondria: signaling, co-compartmentalization, functional coupling, and future directions. Neurochem Res 2020;45:526540.CrossRefGoogle ScholarPubMed
Abramov, AY, Duchen, MR. Mechanisms underlying the loss of mitochondrial membrane potential in glutamate excitotoxicity. Biochim Biophys Acta (BBA)-Bioenergetics. 2008;1777(7–8):953964.CrossRefGoogle ScholarPubMed
Bonora, M, Morganti, C, Morciano, G, et al. Mitochondrial permeability transition involves dissociation of F1 FO ATP synthase dimers and C-ring conformation. EMBO Rep 2017;18(7):10771089.CrossRefGoogle ScholarPubMed
Mnatsakanyan, N, Park, H-A, Wu, J, et al. Mitochondrial ATP synthase c-subunit leak channel triggers cell death upon loss of its F1 subcomplex. Cell Death Differentiation. 2022;29(9):18741887.CrossRefGoogle ScholarPubMed
Li, J, Ma, X, Yu, W, et al. Reperfusion promotes mitochondrial dysfunction following focal cerebral ischemia in rats. PLoS One 2012;7(9):e46498.CrossRefGoogle ScholarPubMed
Liu, D, Wang, H, Zhang, Y, et al. Protective effects of chlorogenic acid on cerebral ischemia/reperfusion injury rats by regulating oxidative stress-related Nrf2 pathway. Drug Design, Dev Therapy. 2020;14:51.CrossRefGoogle ScholarPubMed
Sullivan, PG, Thompson, MB, Scheff, SW. Cyclosporin A attenuates acute mitochondrial dysfunction following traumatic brain injury. Exp Neurology. 1999;160(1):226234.CrossRefGoogle ScholarPubMed
Du, H, Yan, SS. Mitochondrial permeability transition pore in Alzheimer’s disease: cyclophilin D and amyloid beta. Biochim Biophys Acta (BBA)-Molecular Basis Disease. 2010;1802(1):198204.CrossRefGoogle ScholarPubMed
Du, H, Guo, L, Fang, F, et al. Cyclophilin D deficiency attenuates mitochondrial and neuronal perturbation and ameliorates learning and memory in Alzheimer’s disease. Nat Med 2008;14(10):10971105.CrossRefGoogle ScholarPubMed
Ludtmann, MH, Angelova, PR, Horrocks, MH, et al. α-synuclein oligomers interact with ATP synthase and open the permeability transition pore in Parkinson’s disease. Nat Commun 2018;9(1):2293.CrossRefGoogle ScholarPubMed
Kim, H-E, Du, F, Fang, M, et al. Formation of apoptosome is initiated by cytochrome c-induced dATP hydrolysis and subsequent nucleotide exchange on Apaf-1. Proceedings Natl Academy Sciences. 2005;102(49):1754517550.CrossRefGoogle ScholarPubMed
Tatton, NA. Increased caspase 3 and Bax immunoreactivity accompany nuclear GAPDH translocation and neuronal apoptosis in Parkinson’s disease. Exp Neurology. 2000;166(1):2943.CrossRefGoogle ScholarPubMed
Vila, M, Jackson-Lewis, V, Vukosavic, S, et al. Bax ablation prevents dopaminergic neurodegeneration in the 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine mouse model of Parkinson’s disease. Proceedings Natl Academy Sciences. 2001;98(5):28372842.CrossRefGoogle ScholarPubMed
Paradis, E, Douillard, H, Koutroumanis, M, et al. Amyloid β peptide of Alzheimer’s disease downregulates Bcl-2 and upregulates Bax expression in human neurons. J Neurosci 1996;16(23):75337539.CrossRefGoogle ScholarPubMed
Zhang, S, Wang, W. Altered expression of bcl-2 mRNA and Bax in hippocampus with focal cerebral ischemia model in rats. Chin Med Journal. 1999;112(07):608611.Google ScholarPubMed
Park, H-A, Licznerski, P, Mnatsakanyan, N, et al. Inhibition of Bcl-xL prevents pro-death actions of ΔN-Bcl-xL at the mitochondrial inner membrane during glutamate excitotoxicity. Cell Death Differentiation. 2017;24(11):19631974.CrossRefGoogle ScholarPubMed
Lin, MT, Beal, MF. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature. 2006;443(7113):787795.CrossRefGoogle ScholarPubMed
Butterfield, DA, Halliwell, B. Oxidative stress, dysfunctional glucose metabolism and Alzheimer disease. Nat Rev Neurosci 2019;20(3):148160.CrossRefGoogle ScholarPubMed
Su, B, Wang, X, Nunomura, A, et al. Oxidative stress signaling in Alzheimer’s disease. Curr Alzheimer Research. 2008;5(6):525532.CrossRefGoogle ScholarPubMed
Sorbara, MT, Girardin, SE. Mitochondrial ROS fuel the inflammasome. Cell Res 2011;21(4):558560.CrossRefGoogle ScholarPubMed
Wang, W-Y, Tan, M-S, Yu, J-T, et al. Role of pro-inflammatory cytokines released from microglia in Alzheimer’s disease. Ann Translational Medicine. 2015;3(10):136.Google ScholarPubMed
Kawabori, M, Yenari, MA. Inflammatory responses in brain ischemia. Curr Med Chem 2015;22(10):12581277.CrossRefGoogle ScholarPubMed
Chen, C, Turnbull, DM, Reeve, AK. Mitochondrial dysfunction in Parkinson’s disease—cause or consequence? Biology. 2019;8(2):38.CrossRefGoogle ScholarPubMed
Lev, N, Roncevich, D, Ickowicz, D, et al. Role of DJ-1 in Parkinson’s disease. J Mol Neurosci 2006;29:215225.CrossRefGoogle ScholarPubMed
Martin, I, Kim, JW, Dawson, VL, et al. LRRK2 pathobiology in Parkinson’s disease. J Neurochem 2014;131(5):554565.CrossRefGoogle ScholarPubMed
Wang, WJ, Wang, GC, Zhang, M, et al. Isolation of fucoxanthin from the rhizoid of Laminaria japonica Aresch. J Integr Plant Biol. 2005;47(8):10091015.CrossRefGoogle Scholar
Fung, A, Hamid, N, Lu, J. Fucoxanthin content and antioxidant properties of Undaria pinnatifida. Food Chem. 2013;136(2):10551062.CrossRefGoogle ScholarPubMed
Nomura, T, Kikuchi, M, Kubodera, A, et al. Proton-donative antioxidant activity of fucoxanthin with 1, 1-diphenyl-2-picrylhydrazyl (DPPH). IUBMB Life. 1997;42(2):361370.CrossRefGoogle ScholarPubMed
Xia, S, Wang, K, Wan, L, et al. Production, characterization, and antioxidant activity of fucoxanthin from the marine diatom Odontella aurita . Mar Drugs 2013;11(7):26672681.CrossRefGoogle ScholarPubMed
Yan, X, Chuda, Y, Suzuki, M, et al. Fucoxanthin as the major antioxidant in Hijikia fusiformis, a common edible seaweed. Biosci Biotechnol, Biochem 1999;63(3):605607.CrossRefGoogle ScholarPubMed
Sachindra, NM, Sato, E, Maeda, H, et al. Radical scavenging and singlet oxygen quenching activity of marine carotenoid fucoxanthin and its metabolites. J Agric Food Chem. 2007;55(21):85168522.CrossRefGoogle ScholarPubMed
Xiang, S, Liu, F, Lin, J, et al. Fucoxanthin inhibits beta-amyloid assembly and attenuates beta-amyloid oligomer-induced cognitive impairments. J Agric Food Chem. 2017;65(20):40924102.CrossRefGoogle ScholarPubMed
Okubo, H, Miyake, Y, Sasaki, S, et al. Dietary patterns and risk of Parkinson’s disease: a case–control study in Japan. Eur J Neurology. 2012;19(5):681688.CrossRefGoogle ScholarPubMed
Kishida, R, Yamagishi, K, Muraki, I, et al. Frequency of seaweed intake and its association with cardiovascular disease mortality: The JACC Study. J Atherosclerosis Thrombosis. 2020;27(12):13401347.CrossRefGoogle ScholarPubMed
Hashimoto, T, Ozaki, Y, Mizuno, M, et al. Pharmacokinetics of fucoxanthinol in human plasma after the oral administration of kombu extract. Br J Nutr 2012;107(11):15661569.CrossRefGoogle ScholarPubMed
Asai, A, Yonekura, L, Nagao, A. Low bioavailability of dietary epoxyxanthophylls in humans. Br J Nutr 2008;100(2):273277.CrossRefGoogle ScholarPubMed
Hashimoto, T, Ozaki, Y, Taminato, M, et al. The distribution and accumulation of fucoxanthin and its metabolites after oral administration in mice. Br J Nutr 2009;102(2):242248.CrossRefGoogle ScholarPubMed
Manabe, Y, Komatsu, T, Seki, S, et al. Dietary astaxanthin can accumulate in the brain of rats. Biosci Biotechnol, Biochem 2018;82(8):14331436.CrossRefGoogle ScholarPubMed
Vishwanathan, R, Neuringer, M, Snodderly, DM, et al. Macular lutein and zeaxanthin are related to brain lutein and zeaxanthin in primates. Nutr Neuroscience. 2013;16(1):2129.CrossRefGoogle ScholarPubMed
Alzheimer’s Association. Alzheimer’s Disease Facts and Figures. Chicago, IL: Alzheimer’s Association; 2023.Google Scholar
ME, H. Direct evidence of oxidative injury produced by the Alzheimer’s β-amyloid peptide (1–40) in cultured hippocampal neurons. Exp Neurol. 1995;131:193202.Google Scholar
Casley, CS, Canevari, L, Land, JM, et al. Beta-amyloid inhibits integrated mitochondrial respiration and key enzyme activities. J Neurochem. 2002;80(1):91100.CrossRefGoogle ScholarPubMed
Casley, CS, Land, JM, Sharpe, MA, et al. Beta-amyloid fragment 25–35 causes mitochondrial dysfunction in primary cortical neurons. Neurobiol Dis. 2002;10(3):258267.CrossRefGoogle ScholarPubMed
Kuchibhotla, KV, Goldman, ST, Lattarulo, CR, et al. Abeta plaques lead to aberrant regulation of calcium homeostasis in vivo resulting in structural and functional disruption of neuronal networks. Neuron. 2008;59(2):214225.CrossRefGoogle ScholarPubMed
Lin, J, Yu, J, Zhao, J, et al. Fucoxanthin, a marine carotenoid, attenuates beta-amyloid oligomer-induced neurotoxicity possibly via regulating the PI3K/Akt and the ERK pathways in SH-SY5Y cells. Oxid Med Cell Longev. 2017;2017:6792543.CrossRefGoogle ScholarPubMed
Lee, N, Youn, K, Yoon, JH, et al. The role of fucoxanthin as a potent Nrf2 activator via Akt/GSK-3beta/Fyn axis against amyloid-beta peptide-induced oxidative damage. Antioxidants (Basel). 2023;12(3):629.CrossRefGoogle ScholarPubMed
Itoh, K, Wakabayashi, N, Katoh, Y, et al. Keap1 represses nuclear activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes development. 1999;13(1):7686.CrossRefGoogle Scholar
Yang, M, Jin, L, Wu, Z, et al. PLGA-PEG nanoparticles facilitate in vivo anti-Alzheimer’s effects of fucoxanthin, a Marine Carotenoid derived from edible brown algae. J Agric Food Chem. 2021;69(34):97649777.CrossRefGoogle ScholarPubMed
Kawee-ai, A, Kuntiya, A, Kim, SM. Anticholinesterase and antioxidant activities of fucoxanthin purified from the microalga Phaeodactylum tricornutum . Nat Prod Commun 2013;8(10):13811386.Google ScholarPubMed
Dickson, DW. Parkinson’s disease and parkinsonism: neuropathology. Cold Spring Harbor Perspect Medicine. 2012;2(8):a009258.CrossRefGoogle Scholar
Jung, HA, Ali, MY, Choi, RJ, et al. Kinetics and molecular docking studies of fucosterol and fucoxanthin, BACE1 inhibitors from brown algae Undaria pinnatifida and Ecklonia stolonifera . Food Chem Toxicol 2016;89:104111.CrossRefGoogle ScholarPubMed
WHO. Parkinson Disease. Geneva: WHO; 2023.Google Scholar
Parkinson’s Foundation. Understanding Parkinson’s. Miami, FL: Parkinson’s Foundation; 2023.Google Scholar
Tambasco, N, Romoli, M, Calabresi, P. Levodopa in Parkinson’s disease: current status and future developments. Curr Neuropharmacol 2018;16(8):12391252.CrossRefGoogle ScholarPubMed
Grünewald, A, Rygiel, KA, Hepplewhite, PD, et al. Mitochondrial DNA depletion in respiratory chain–deficient Parkinson disease neurons. Ann Neurol 2016;79(3):366378.CrossRefGoogle ScholarPubMed
Moon, HE, Paek, SH. Mitochondrial dysfunction in Parkinson’s disease. Exp Neurobiology. 2015;24(2):103.CrossRefGoogle ScholarPubMed
Pickrell, AM, Youle, RJ. The roles of PINK1, parkin, and mitochondrial fidelity in Parkinson’s disease. Neuron. 2015;85(2):257273.CrossRefGoogle ScholarPubMed
Chen, R, Park, H-A, Mnatsakanyan, N, et al. Parkinson’s disease protein DJ-1 regulates ATP synthase protein components to increase neuronal process outgrowth. Cell death disease. 2019;10(6):112.CrossRefGoogle ScholarPubMed
Nakamura, K, Nemani, VM, Azarbal, F, et al. Direct membrane association drives mitochondrial fission by the Parkinson disease-associated protein α-synuclein*♦. J Biol Chem 2011;286(23):2071020726.CrossRefGoogle ScholarPubMed
Thorne, NJ, Tumbarello, DA. The relationship of alpha-synuclein to mitochondrial dynamics and quality control. Front Mol Neurosci 2022;15:947191.CrossRefGoogle ScholarPubMed
Wu, W, Han, H, Liu, J, et al. Fucoxanthin prevents 6-OHDA-induced neurotoxicity by targeting Keap1. Oxid Med Cell Longev. 2021;2021:6688708.Google ScholarPubMed
Liu, J, Lu, Y, Tang, M, et al. Fucoxanthin prevents long-term administration l-DOPA-induced neurotoxicity through the ERK/JNK-c-Jun system in 6-OHDA-lesioned mice and PC12 cells. Mar Drugs. 2022;20(4):245.CrossRefGoogle ScholarPubMed
Lian, W, Hu, X, Zhang, J, et al. Fucoxanthin protects retinal ganglion cells and promotes parkin-mediated mitophagy against glutamate excitotoxicity. Neuroreport. 2023;34(7):385.CrossRefGoogle ScholarPubMed
Sims, NR, Mitochondria, Muyderman H., oxidative metabolism and cell death in stroke. Biochim Biophys Acta (BBA)-Molecular Basis Disease. 2010;1802(1):8091.CrossRefGoogle ScholarPubMed
Yang, J-L, Mukda, S, Chen, S-D. Diverse roles of mitochondria in ischemic stroke. Redox Biol 2018;16:263275.CrossRefGoogle ScholarPubMed
Broughton, BR, Reutens, DC, Sobey, CG. Apoptotic mechanisms after cerebral ischemia. Stroke. 2009;40(5):e3319.CrossRefGoogle ScholarPubMed
Li, H, Park, H-A, Jonas, EA. Fluorescent measurement of synaptic activity using synaptophluorin in isolated hippocampal neurons. Bio-Protocol. 2014;4(23):e1304-e.CrossRefGoogle ScholarPubMed
Baumgartner, HK, Gerasimenko, JV, Thorne, C, et al. Calcium elevation in mitochondria is the main Ca2+ requirement for mitochondrial permeability transition pore (mPTP) opening. J Biol Chem 2009;284(31):2079620803.CrossRefGoogle ScholarPubMed
Ikeda, K, Kitamura, A, Machida, H, et al. Effect of Undaria pinnatifida (Wakame) on the development of cerebrovascular diseases in stroke-prone spontaneously hypertensive rats. Clin Exp Pharmacol Physiol 2003;30(1-2):4448.CrossRefGoogle ScholarPubMed
Hu, L, Chen, W, Tian, F, et al. Neuroprotective role of fucoxanthin against cerebral ischemic/reperfusion injury through activation of Nrf2/HO-1 signaling. Biomed Pharmacother 2018;106:14841489.CrossRefGoogle ScholarPubMed
Wang, Q, Zhang, D, Lu, J, et al. PLGA-PEG-fucoxanthin nanoparticles protect against ischemic stroke in vivo. J Funct Foods 2022;99:105359.CrossRefGoogle Scholar
Kalin, NH. The critical relationship between anxiety and depression. Am Psychiatric Assoc. 2020;177(5):365367.CrossRefGoogle ScholarPubMed
Fox, ME, Lobo, MK. The molecular and cellular mechanisms of depression: a focus on reward circuitry. Mol Psychiatry 2019;24(12):17981815.CrossRefGoogle ScholarPubMed
Hollis, F, Pope, BS, Gorman-Sandler, E, et al. Neuroinflammation and mitochondrial dysfunction link social stress to depression. Neurosci Social Stress. 2022:5993.CrossRefGoogle ScholarPubMed
Li, C-T, Su, T-P, Wang, S-J, et al. Prefrontal glucose metabolism in medication-resistant major depression. Br J Psychiatry. 2015;206(4):316323.CrossRefGoogle ScholarPubMed
Gong, Q, Yan, X-J, Lei, F, et al. Proteomic profiling of the neurons in mice with depressive-like behavior induced by corticosterone and the regulation of berberine: pivotal sites of oxidative phosphorylation. Mol Brain. 2019;12:114.CrossRefGoogle ScholarPubMed
Lin, S, Huang, L, Luo, Z-C, et al. The ATP level in the medial prefrontal cortex regulates depressive-like behavior via the medial prefrontal cortex-lateral habenula pathway. Biol Psychiatry 2022;92(3):179192.CrossRefGoogle ScholarPubMed
Kosten, TA, Galloway, MP, Duman, RS, et al. Repeated unpredictable stress and antidepressants differentially regulate expression of the bcl-2 family of apoptotic genes in rat cortical, hippocampal, and limbic brain structures. Neuropsychopharmacology. 2008;33(7):15451558.CrossRefGoogle ScholarPubMed
Wang, X, Xie, Y, Zhang, T, et al. Resveratrol reverses chronic restraint stress-induced depression-like behaviour: involvement of BDNF level, ERK phosphorylation and expression of Bcl-2 and Bax in rats. Brain Res Bull 2016;125:134143.CrossRefGoogle ScholarPubMed
Kris-Etherton, PM, Petersen, KS, Hibbeln, JR, et al. Nutrition and behavioral health disorders: depression and anxiety. Nutr Reviews. 2021;79(3):247260.CrossRefGoogle ScholarPubMed
Molendijk, M, Molero, P, Sánchez-Pedreño, FO, et al. Diet quality and depression risk: a systematic review and dose-response meta-analysis of prospective studies. J Affective Disorders. 2018;226:346354.CrossRefGoogle Scholar
Guo, F, Huang, C, Cui, Y, et al. Dietary seaweed intake and depressive symptoms in Japanese adults: a prospective cohort study. Nutr J 2019;18(1):18.CrossRefGoogle ScholarPubMed
Park, I, Kim, J, Kim, M, et al. Sargassum horneri extract attenuates depressive-like behaviors in mice treated with stress hormone. Antioxidants. 2023;12(10):1841.CrossRefGoogle ScholarPubMed
Chao, CC, Ma, YL, Lee, EH. Brain-derived neurotrophic factor enhances Bcl-xL expression through protein kinase casein kinase 2-activated and nuclear factor kappa B-mediated pathway in rat hippocampus. Brain Pathology. 2011;21(2):150162.CrossRefGoogle ScholarPubMed
Park, H-A, Crowe-White, KM, Ciesla, L, et al. Alpha-tocotrienol enhances arborization of primary hippocampal neurons via upregulation of Bcl-xL. Nutr Res 2022;101:3142.CrossRefGoogle ScholarPubMed
Jiang, X, Wang, G, Lin, Q, et al. Fucoxanthin prevents lipopolysaccharide-induced depressive-like behavior in mice via AMPK-NF-κB pathway. Metab Brain Disease. 2019;34:431442.CrossRefGoogle ScholarPubMed
Harman, D. Free radical theory of aging. Mutat Research/DNAging. 1992;275(3–6):257266.CrossRefGoogle ScholarPubMed
Paradies, G, Petrosillo, G, Paradies, V, et al. Mitochondrial dysfunction in brain aging: role of oxidative stress and cardiolipin. Neurochem Int 2011;58(4):447457.CrossRefGoogle ScholarPubMed
Venkateshappa, C, Harish, G, Mahadevan, A, et al. Elevated oxidative stress and decreased antioxidant function in the human hippocampus and frontal cortex with increasing age: implications for neurodegeneration in Alzheimer’s disease. Neurochem Res 2012;37:16011614.CrossRefGoogle ScholarPubMed
Boveris, A, Navarro, A. Brain mitochondrial dysfunction in aging. IUBMB Life. 2008;60(5):308314.CrossRefGoogle ScholarPubMed
Willcox, BJ, Willcox, DC. Caloric restriction, CR mimetics, and healthy aging in Okinawa: controversies and clinical implications. Curr Opin Clin Nutr Metab Care. 2014;17(1):51.Google Scholar
Willcox, BJ, Willcox, DC, Todoriki, H, et al. Caloric restriction, the traditional Okinawan diet, and healthy aging: the diet of the world’s longest-lived people and its potential impact on morbidity and life span. Ann NY Acad Sci 2007;1114(1):434455.CrossRefGoogle ScholarPubMed
Baierle, M, Nascimento, SN, Moro, AM, et al. Relationship between inflammation and oxidative stress and cognitive decline in the institutionalized elderly. Oxid Med Cell Longevity 2015;2015:804198.CrossRefGoogle ScholarPubMed
Ownby, RL. Neuroinflammation and cognitive aging. Curr Psychiatry Reports. 2010;12:3945.CrossRefGoogle ScholarPubMed
Chen, Y, Dong, J, Gong, L, et al. Fucoxanthin, a marine derived carotenoid, attenuates surgery-induced cognitive impairments via activating Akt and ERK pathways in aged mice. Phytomedicine. 2023;120:155043.CrossRefGoogle ScholarPubMed
Lashmanova, E, Proshkina, E, Zhikrivetskaya, S, et al. Fucoxanthin increases lifespan of Drosophila melanogaster and Caenorhabditis elegans. Pharmacol Res 2015;100:228241.CrossRefGoogle ScholarPubMed
Moskalev, A, Shaposhnikov, M, Zemskaya, N, et al. Transcriptome analysis reveals mechanisms of geroprotective effects of fucoxanthin in Drosophila . BMC Genomics 2018;19(3):6576.CrossRefGoogle ScholarPubMed
Dagda, RK. Role of mitochondrial dysfunction in degenerative brain diseases, an overview. Brain Sciences. 2018;8(10):178.CrossRefGoogle ScholarPubMed
Figure 0

Table 1. The effects of fucoxanthin on in vitro and in vivo brain disorder models

Figure 1

Fig. 1. Mechanism of mitochondrial protection by fucoxanthin in brain disorders.