Next Article in Journal
A Metabolomics and Big Data Approach to Cannabis Authenticity (Authentomics)
Next Article in Special Issue
Chronic Liver Disease: Latest Research in Pathogenesis, Detection and Treatment
Previous Article in Journal
Unraveling the Host Genetic Background Effect on Internal Organ Weight Influenced by Obesity and Diabetes Using Collaborative Cross Mice
Previous Article in Special Issue
Detection of Novel Biomarkers in Pediatric Autoimmune Hepatitis by Proteomic Profiling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Fucoxanthin in Non-Alcoholic Fatty Liver Disease

1
Division of Bio-Medical Science and Technology, KIST School, University of Science and Technology, Seoul 02792, Republic of Korea
2
Natural Product Informatics Research Center, KIST Gangneung Institute of Natural Products, Gangneung 25451, Republic of Korea
3
Microalgae Ask US Co., Ltd., Gangneung 25441, Republic of Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(9), 8203; https://doi.org/10.3390/ijms24098203
Submission received: 20 March 2023 / Revised: 27 April 2023 / Accepted: 1 May 2023 / Published: 3 May 2023

Abstract

:
Chronic liver disease (CLD) has emerged as a leading cause of human deaths. It caused 1.32 million deaths in 2017, which affected men more than women by a two-to-one ratio. There are various causes of CLD, including obesity, excessive alcohol consumption, and viral infection. Among them, non-alcoholic fatty liver disease (NAFLD), one of obesity-induced liver diseases, is the major cause, representing the cause of more than 50% of cases. Fucoxanthin, a carotenoid mainly found in brown seaweed, exhibits various biological activities against NAFLD. Its role in NAFLD appears in several mechanisms, such as inducing thermogenesis in mitochondrial homeostasis, altering lipid metabolism, and promoting anti-inflammatory and anti-oxidant activities. The corresponding altered signaling pathways are the β3-adorenarine receptor (β3Ad), proliferator-activated receptor gamma coactivator (PGC-1), adenosine monophosphate-activated protein kinase (AMPK), peroxisome proliferator-activated receptor (PPAR), sterol regulatory element binding protein (SREBP), nuclear factor kappa B (NF-κB), mitogen-activated protein kinase (MAPK), protein kinase B (AKT), SMAD2/3, and P13K/Akt pathways. Fucoxanthin also exhibits anti-fibrogenic activity that prevents non-alcoholic steatohepatitis (NASH) development.

Graphical Abstract

1. Introduction

Chronic liver disease (CLD) is the leading disease resulting in death in the world. There were approximately 1.5 billion cases of CLD worldwide in 2020, with 1.32 million deaths in 2017, which affected men more than women by a two-to-one ratio [1,2]. Originally, viral hepatitis was the major etiology of CLD, but recently, obesity and alcohol consumption have become increasingly common factors for CLD. CLD caused by a viral infection can be treated with proper medication and prevented with vaccination. It can also be easily and precisely detected based on the tests for infection, unlike non-alcoholic liver disease [3]. Non-alcoholic fatty liver disease (NAFLD) occurs when fat accumulates excessively in the liver due to an obesogenic diet and lifestyle [4]; hence, diagnosing NAFLD is challenging. The metabolic dysregulation caused by obesity disrupts the body’s homeostasis and affects the function of the liver. Prevention and treatment of obesity are thus critically important to combatting and preventing NAFLD.
Fucoxanthin, a naturally derived carotenoid compound, has shown anti-obesity effects [5,6,7,8]. It was first isolated from brown algae and seaweed such as Fucus, Dictyota, and Laminaria by Willstätter and Page in 1914 [9]. Brown seaweeds are a common dietary food in many parts of Asia and are found naturally in open seas, where they are exposed to metals as well as metalloids [10]. It is now commercially produced by various technologies [11], such as bioreactors using microalgae and diatoms. As a natural compound, it is hypothesized to have less severe side effects with lower toxicity levels. Fucoxanthin shows various kind of bioactivities, including anti-obesity effects that can be used for the treatment and prevention of NAFLD. Here, in order to better understand fucoxanthin’s activity on NAFLD, we review the literature on the activity of fucoxanthin in NAFLD and its signaling pathways.

2. Structure and Metabolites of Fucoxanthin

Fucoxanthin is a carotenoid mainly found in brown seaweed, acting as a photosynthetic pigment along with chlorophyll a and c, and β-carotene [12]. With a molecular weight of 658.9 g/mol, fucoxanthin (C42H58O6) is located in the photosynthetic organ of brown seaweed and microalgae and is responsible for photochemical events [13,14]. It has a unique structure with a wide range of inherent activities. There is an allenic bond, a conjugated carbonyl, a 5,6-monoepoxide, and an acetyl group [15,16]. An allenic bond is a condition where one carbon atom has two double bonds with each adjacent carbon [17].
Due to the structure of fucoxanthin, it mainly has two biological functions: singlet oxygen quenching and free radical scavenging [18]. The singlet oxygen quenching activity is due to the carbon double bonds located in the backbone of fucoxanthin. It mainly depends on physical quenching without any chemical reaction [19]. It works by transferring the energy in the singlet oxygen molecules to the conjugated double bond in fucoxanthin [20,21]. The excited fucoxanthin can dissipate energy into the environment, returning it to a ground state [19]. That will bring the fucoxanthin molecule back to its original state. Singlet oxygen quenching activity depends on the number of conjugated carbon double bonds [22,23]. The higher the number of conjugated double bonds, the more energy that can be transferred from the singlet oxygen molecule to the carotenoid.
Meanwhile, the free radical scavenging ability of fucoxanthin is due to the functional groups in the terminal rings [19,24]. They function as electron acceptors and electron donors, as well as in adduct formation. Moreover, there is an allenic bond in fucoxanthin. Among the 700 carotenoids in nature, there are about 40 types of carotenoids, which have an allenic bond, including fucoxanthin [25]. The allenic bond provides carotenoids with a higher activity than alkenes and a peculiar axial chirality [26] that contributes to the activity of fucoxanthin.
Fucoxanthin can be absorbed into the human body through the digestive system at the intestinal level. Fucoxanthin is metabolized in the liver via fucoxanthinol to amarouciaxanthin A (Figure 1), requiring the cofactor nicotinamide adenine dinucleotide phosphate (NADP) [27,28,29]. Several enzymes are also involved in the gastrointestinal tract, such as lipase and cholesterol esterase. It was reported that the proportions of fucoxanthin, fucoxanthinol, and amarouciaxanthin A in the adipose tissue were 13%, 32%, and 55%, respectively, whereas in other tissues, including the liver, lungs, kidney, heart, and spleen, were 1–11%, 63–76%, and 20–26%, respectively [6,30]. Other than fucoxanthinol and amarouciaxanthin A, another fucoxanthin metabolite derived from fucoxanthinol, halocynthiaxanthin, has been isolated from Undaria pinnatifida [12]. This metabolite has not yet been fully studied, and discovering other metabolites of fucoxanthin is possible.
Fucoxanthin toxicity has been tested through several experiments, including in animals and humans. In ICR mice, fucoxanthin showed no mortality and no abnormalities in a single-dose study at 1000 and 2000 mg/kg, as well as 500 and 1000 mg/kg in a repeated-dose study for 30 days [14,31]. A single oral dose study was also performed on rats and showed no toxicity with a fucoxanthin administration of 200 mg/kg body weight [32]. It is also declared safe at 0.5% w/v for application on human skin [33]. Hence, the administration of fucoxanthin has been determined to be safe, and the physiological activity of fucoxanthin is considered to have great potential.

3. Role of Fucoxanthin in Non-Alcoholic Fatty Liver Disease

NAFLD occurs when the rate of hepatic fatty acid uptake is greater than its oxidation [34]. NAFLD is closely associated with obesity and the increase in intrahepatic triglyceride (IHTG) [35]. This occurs with an imbalance in food intake and energy expenditure, resulting in fat accumulation [36]. Triglycerides accumulate in the liver and disrupt the body’s metabolism, from the normal state to the hypercaloric state, when the energy produced inside the body is more than sufficient; hence, it is stored as a lipid [37]. Obesity increases mortality in NAFLD patients [38]. The rising number of NAFLD cases is closely related to the rising trend of obesity [39]. It is presumed that combating obesity is important for the treatment and prevention of NAFLD.
Fundamentally, there are two ways to overcome obesity: increasing energy expenditure or decreasing energy gain by controlling food intake [40]. Either way, the system works by adjusting the hypercaloric metabolic state of the body back to the homeostatic state. Fucoxanthin is known to have an anti-obesity effect, proven through different kinds of experiments, including cell culture, animal models, and human studies. The anti-obesity mechanisms of fucoxanthin are categorized into two classes: inducing thermogenic activity and altering lipid metabolism. Altering lipid metabolism is a strategy that can be effectively used against NAFLD [41].
Fucoxanthin has also been reported to reduce hepatic injury by decreasing hepatic fat accumulation and liver weight gain in a choline-deficient, L-amino-acid-defined high-fat diet (CDAHFD), non-alcoholic steatohepatitis (NASH) mouse model. It decreased hepatic lipid oxidation and NASH inflammation by inhibiting the production of chemokines [42]. It also alleviates lipid peroxidation in hepatocytes, resulting in the suppression of lipid accumulation [43]. Fucoxanthinol and amarouciaxanthin A have also been found to have an anti-inflammatory effect against NASH by down-regulating the hepatic stellate cell marker [42]. In conclusion, fucoxanthin is a promising therapeutic for inhibiting hepatic inflammation and preventing fibrosis in liver disease. The role of fucoxanthin in liver diseases, especially NAFLD, can be explained through multiple mechanisms, such as thermogenesis-induced anti-obesity activity, altered lipid metabolism, and anti-inflammatory, anti-oxidant, and anti-fibrogenic activities (Figure 2).

3.1. Fucoxanthin Affects Mitochondrial Homeostasis through Thermogenic Activity

Mitochondria play a major role in human health. Their role is centered on homeostasis and energy metabolism, including maintaining and producing the energy needed by the human body [4]. The disruption of mitochondrial homeostasis and elevated oxidative stress are commonly observed in fatty liver disease patients [44] and are characterized by a reduction in respiratory chain activity and impaired mitochondrial β-oxidation [45].
Many biological activities are performed inside mitochondria. One of them is oxidative phosphorylation. Mammalian cells synthesize energy through oxidizing substrates in inside mitochondria. One of the chemical reactions involved is oxidative phosphorylation located in mitochondria. During oxidative phosphorylation, free energy is converted into the displacement of adenosine triphosphate (ATP) in an equilibrium reaction. However, with uncoupling protein, hydrogen and heat are released instead of ATP synthesis through the uncoupling to ATP synthase, bypassing ATP synthase, and is hence called an uncoupling protein [46,47,48], as shown in Figure 3. Proton leak (hydrogen leak) in mitochondria is expected to occur through electron escape from mitochondrial oxidoreductase to generate superoxide [49].
Fucoxanthin exhibits anti-obesity effects, mainly through thermogenic effects via mitochondrial uncoupling protein 1 (UCP1) [5,8], as shown in Figure 3. UCP1, a 32 kDa protein, is an inner mitochondrial membrane protein that is a molecular basis for the protonophore activity in the mitochondrial inner membrane. UCP-1 allows protons to enter the mitochondrial matrix at a lower energy with the proton leak [48]. Fucoxanthin induces the expression of UCP1 in abdominal white adipose tissue (WAT) [51]. This phenomenon is also known as browning, where WAT changes in phenotype into brown adipose tissue (BAT) [52]. BAT plays a role in dissipating energy through heat production, unlike WAT, which stores excess energy as triglycerides [52,53,54]. As the browning process occurs, energy expenditure in WAT is upregulated [55]. Through this process, the amount of WAT is reduced, which means fat accumulation is also reduced. This mechanism helps to treat NAFLD.
Fucoxanthin induces loss of WAT and alleviates hyperglycemia in an obese–diabetic KK-A(y) mouse model [56]. The same effect was also shown in high-fat-diet (HFD)-induced obese mice, as well as hyperinsulinemia and hyperleptinemia effects [57]. The decrease in WAT is consistently followed by an increase in BAT weight [58,59,60]. Further, the expression of UCP-1 is also increased in the WAT of KK-A(y) mice fed with fucoxanthin [61]. This anti-obesity activity of fucoxanthin was also examined in a human study, where according to Mikami (2017), fucoxanthin reduced HbA1c levels in subjects with G/G alleles of the UCP1 gene compared to those with the A/A and A/G alleles (thrifty allele of UCP1-3826A/G) in a human study consisting of 60 normal weight and obese Japanese adults with a BMI over 22 [62]. By lowering the level of HbA1c, fucoxanthin supplementation succeeded in lowering blood sugar levels, with no significant effect on visceral fat.
UCP1 is related to other mitochondrial metabolite transporters, such as the adenine nucleotide translocator, a proton channel in the mitochondrial inner membrane that permits the translocation of protons from the mitochondrial intermembrane space to the mitochondrial matrix. Other than the upregulation of UCP1, the expression of the β3-adrenergic receptor (β3Ad) and peroxisome proliferator-activated receptor gamma coactivator 1 (PGC-1) is also upregulated in WAT [57,63]. Activation of PGC-1 induces mitochondrial biogenesis [62]. Fucoxanthin is presumed to induce mitochondrial biogenesis.

3.2. Fucoxanthin Alters Lipid Metabolism

Lipid metabolism can be elucidated in two ways, lipolysis and lipogenesis. Both work in opposing manners, where lipogenesis synthesizes fat and lipolysis breaks down fat. Fat accumulation, which leads to obesity, results when an imbalance in lipogenesis and lipolysis occurs [64]. One of NAFLD’s characteristics is an alteration in the lipid metabolism that is also observed in atherogenic dyslipidemia [65]. An increase in de novo lipogenesis is considered the major alteration of lipid metabolism in NAFLD [66]. As a result, hepatic steatosis, in which the intrahepatic fat of more than 5% of the liver’s weight is accumulated, occurs through an increase in liver fatty acids and downregulation of β-oxidation [67].
Other than inducing thermogenesis activity via UCP-1 activation, fucoxanthin also functions by altering lipid metabolism and absorption [68], as shown in Figure 4 by inhibiting lipogenesis while promoting lipolysis. This mechanism works against NAFLD-altered lipid metabolism. Fucoxanthin upregulates enzymes related to lipolysis while downregulating enzymes related to lipogenesis. In HFD-fed mice, fucoxanthin can decrease hepatic lipid and plasma triacylglycerol levels [7,69]. These effects were shown through the increase in undigested fecal lipids. Fucoxanthin also functions by reducing the activity of hepatic lipogenesis and upregulating the activity of fatty acid β-oxidation [7]. It upregulates other key proteins in lipid metabolism, such as AMP-activated protein kinase and acetyl-CoA carboxylase in epididymal adipose tissue. It also induces β3Ad [51], which upregulates lipolysis and thermogenesis [70].
The supplementation of fucoxanthin in an obese mouse model induced through HFD (20% fat) decreased the visceral fat pads without altering the food intake [69]. It also increased adiponectin levels and, on the other hand, decreased leptin levels in plasma. Both adiponectin and leptin are adipose-derived hormones that act as messengers to deliver signals from the adipose tissues to other tissues and organs [71]. Adiponectin has been reported to act as an insulin-sensitizing adipokine in heterozygous peroxisome proliferator-activated receptor (PPAR) β knockout mice, which protects it from HFD-induced obesity [72]. PPAR γ is also reported to be downregulated in 3T3-L1 adipocytes with fucoxanthin supplementation [8]. Fucoxanthin supplementation through Nitzschia laevis extract (NLE) has also been reported to decrease abdominal fat and hepatic steatosis in C57BL/6J mice and to avert the accumulation of lipids in HepG2 cells. It elevates mitochondrial activity, shown through the enhanced oxygen consumption rate and mitochondrial membrane potential, and phosphorylated acetyl-CoA carboxylase [67]. The phosphorylation of acetyl-CoA carboxylase inhibits lipogenesis in the hepatocytes of rats and prevents the accumulation of hepatic lipids [73].
Standardized fucoxanthin extract from Phaeodactylum tricornutum showed inhibitory activity toward lipogenesis in 3T3-L1 adipocytes by decreasing intracellular lipid contents without any cytotoxicity [8]. Another fucoxanthin extract from Petalonia binghamiae has been reported to suppress the accumulation of lipid droplets in the liver as well as alanine and aspartate transaminase serum levels. It functions by upregulating the adenosine monophosphate-activated protein kinase (AMPK) signaling pathway, targeting acetyl-CoA carboxylase in adipocytes as well as fatty acid β-oxidation [74]. The fatty acid synthase (FAS) protein, which is also included in the AMPK signaling pathway, is involved in lipogenesis and has also been reported to be downregulated in the livers of db/db diabetic mice [19,75]. Further, the expression of PPAR α, phosphorylated acetyl-CoA carboxylase, and carnitine palmitoyltransferase 1 was upregulated. Similarly, the fucoxanthin extracted from Undaria pinnatifida also exhibits an anti-obesity effect in the HFD mouse model, which was demonstrated by decreased visceral fat and hepatic lipid accumulation, as well as decreased adipocyte size [9,76].
Fucoxanthin as a medication for NAFLD has recently reached clinical trials. The administration of fucoxanthin combined with fucoidan for 6 months in 21 patients with NAFLD attenuated hepatic lipotoxicity. The fucoxanthin–fucoidan treatment succeeded in lowering triglyceride, total cholesterol, alanine transaminase (ALT), as well as aspartate aminotransferase (AST) in a high fat diet mouse model administered 200 or 400 mg/kg bw fucoidan–fucoxanthin. It also significantly reduced the NAFLD-induced inflammatory cytokines IL-6 and IFN-γ. Leptin and adiponectin were also altered favorably to fight against NAFLD [77]. However, fucoxanthin at low concentrations, 0.015% and 0.03% w/w, did not effectively reduce triglycerides and total cholesterol in the high fat diet mouse model [78]. Another study with the same high fat diet mouse model accompanied by fucoxanthin supplementation at 0.05% or 0.2% w/w successfully decreased triglyceride and cholesterol levels [76]. Hence, the effects of fucoxanthin on altering lipid metabolism, specifically in the high fat diet mouse model, are dependent on dose and feed composition. Further studies should be conducted in order to determine the minimum effective dose of fucoxanthin in this model.

3.3. Anti-Inflammatory Activity of Fucoxanthin

NAFLD has a broad spectrum of characteristics, including inflammation. Moreover, obesity is also categorized as low-level inflammation marked with abnormally produced inflammatory adipocytokines [79]. Inflammation in NAFLD is triggered by an excessive accumulation of lipids, which activates hepatic fibrosis [80]. NAFLD progression is affected by the balance between pro- and anti-inflammatory stimuli. Once inflammatory cells in the liver are activated, steatosis occurs, and inflammation develops. An increase in fatty acids in the liver will also increase systemic inflammation, which is commonly observed in NAFLD patients [81].
In NAFLD, chronic inflammation has also been proposed to be promoted by insulin resistance [82]. The inflammation in NAFLD is characterized by the upregulation of several cytokines, including IL-6, IL-1β, and TNF-α. This fact was proven by an in vivo study with the NAFLD mouse model, which resulted in elevated TNF-α levels in the liver [83]. Another in vivo study using the NAFLD-HFD mouse model exhibited similar results, where AST, ALP, leptin, cholesterol, triglyceride, TNF-α, and TGF-β levels were significantly increased [84]. It can be concluded that pro-inflammatory cytokines are upregulated in NAFLD.
Fucoxanthin exhibits anti-inflammatory activity, as proven by the various studies indicated in Table 1. Fucoxanthin can alter PPAR signaling pathways. It is well known that the PPAR signaling pathways are related not only to lipid metabolism but also inflammation. They control inflammation and the immune response by regulating macrophages [85]. Therefore, modulation of the PPAR signaling pathway for anti-obesity effects will simultaneously exhibit an anti-inflammatory effect. This theory was proven by several studies focusing on the anti-inflammatory effect of fucoxanthin. Fucoxanthin supplementation successfully suppressed IL-1β, TNF-α, COX-2, and iNOS in an HFD mouse model [79]. It also suppressed pro-inflammatory factors such as NO, PGE2, IL-1β, TNF-α, and IL-6 in RAW 264.7 cells via the NF-κB and MAPK signaling pathways [86]. Fucoxanthin supplementation in a diabetic/obese KK-Ay mouse model succeeded in downregulating the expression of inflammatory cytokines such as TNF-α, IL-6, and monocyte chemoattractant protein-1 (Mcp-1). Fucoxanthin also inhibited macrophage infiltration into the white adipose tissues of the mice [43].
An in vitro study using Raw264.7 macrophages successfully showed the downregulation of IL-10, IL-6, iNOS, COX-2, and NF-κB signals [87]. In vivo studies using the LPS-induced sepsis mouse model also exhibited the same results, where inflammatory cytokines such as IL-6, IL-1β, and TNF-α were downregulated after fucoxanthin supplementation. Fucoxanthin is presumed to inhibit the phosphorylation of the NF-κB signaling pathway at the cellular level and block nuclear translocation [88]. Fucoxanthin also downregulated iNOS and COX-2 expression in a carrageenan-induced paw edema mouse experiment through the MAPK, Akt, and NF-κB signaling pathways [89]. Other interleukins, such as IL-4, IL-5, IL-8, and IL-13, were also reported to be downregulated in asthmatic mouse models [90].
Table 1. Inflammatory cytokines altered after fucoxanthin supplementation.
Table 1. Inflammatory cytokines altered after fucoxanthin supplementation.
Inflammatory CytokinesExperimental ModelDoseReferences
↓ IL-1βHigh-fat-diet-induced obese mice0.2, 0.4, or 0.6%[79]
RAW 264.7 macrophagesI. okamurae-extracted fucoxanthin; 12.5, 25, or 50 μM[86]
RAW 264.7 macrophages5, 10, or 20 μM[87]
↓ IL-4
↓ IL-5
OVA-stimulated (OVA) mice10 or 30 μM[90]
Bronchoalveolar lavage fluid (BALF) from asthmatic mice10 mg/kg or 30 mg/kg
OVA-stimulated (OVA) mice10 or 30 μM
↓ IL-6RAW 264.7 macrophages cellsI. okamurae-extracted fucoxanthin; 12.5, 25, or 50 μM[86]
BEAS-2B cells3, 10, or 30 μM[90]
Bronchoalveolar lavage fluid (BALF) from asthmatic mice10 mg/kg or 30 mg/kg
RAW 264.7 macrophages5, 10, or 20 μM[87]
↓ IL-8BEAS-2B cells3, 10, or 30 μM[90]
Bronchoalveolar lavage fluid (BALF) from asthmatic mice10 mg/kg or 30 mg/kg
↓ IL-10RAW 264.7 macrophages5, 10, or 20 μM[87]
↓ IL-13Bronchoalveolar lavage fluid (BALF) from asthmatic mice10 mg/kg or 30 mg/kg[90]
OVA-stimulated (OVA) mice10 or 30 μM
↓ TNF-αHigh-fat-diet-induced obese mice0.2, 0.4, 0.6%[79]
RAW 264.7 macrophages cellsI. okamurae-extracted fucoxanthin; 12.5, 25, or 50 μM[86]
Bronchoalveolar lavage fluid (BALF) from asthmatic mice10 mg/kg or 30 mg/kg[90]
OVA-stimulated (OVA) mice10 or 30 μM
↓ COX-2High-fat-diet-induced obese mice0.2, 0.4, or 0.6%[79]
RAW 264.7 macrophages5, 10, or 20 μM[87]
↓ iNOS
↑—upregulated; ↓—downregulated.

3.4. Anti-Oxidant Activity of Fucoxanthin against NAFLD

Inflammation is correlated with reactive oxygen species (ROS), where increased ROS promote inflammation [91,92]. The increase in oxidative stress caused by excessive ROS production is one of the causes of NAFLD. Excess production of ROS has also been shown in patients with type 2 diabetes, insulin resistance, obesity, NASH, and NAFLD [81]. ROS are highly reactive and unstable, so they often lead to an imbalance in the bioavailability of the cellular anti-oxidant system [93]. When that occurs, it disrupts intracellular metabolism and modifies the functional role of cellular enzymes, structural proteins, and even cell membranes [94].
Oxidative stress may also occur as a form of lipid-induced stress. Lipid accumulation in the liver may induce oxidative stress by altering mitochondrial activity and function and disrupting the anti-oxidant system. ROS are involved in the mitochondrial respiratory chain. In response to fatty acid oxidation disruption, which takes place in mitochondria, the production of ROS may be increased. This may also lead to cell apoptosis; hence, oxidative stress may be a factor in steatosis inflammatory progression [81]. In NAFLD, the excess ROS affect lipid peroxidation and impair mitochondrial and peroxisomal oxidation of fatty acids, resulting in the release of inflammatory cytokines [95], as well as the production of pro-inflammatory cytokines via MAPK phosphorylation [91].
Although fucoxanthin lacks pro-vitamin A activity, it has unique properties as a carotenoid that exhibits anti-oxidant activity [96]. Through its anti-oxidant activity, it exhibits anti-inflammatory effects related to obesity. ROS formation was reduced when PC12 cells were treated with fucoxanthin [79]. Another study using 3T3-L1 cells showed a decrease in enzymes related to the production of ROS, such as nicotinamide adenine dinucleotide phosphate (NADPH) oxidase 4 (NOX4), the NADPH-generating enzyme, and glucose-6-phosphate dehydrogenase after fucoxanthin treatment [97]. Fucoxanthin also exhibits free radical scavenging activity, as shown by several studies. Fucoxanthin treatment succeeded in reducing doxorubicin-induced ROS compared to primary cardiomyocytes treated with doxorubicin alone, indicating that the anti-oxidant effect of fucoxanthin exerts a cardioprotective effect [98]. Fucoxanthin supplementation in a diabetic/obese KK-Ay mouse model reduced oxidative stress by alleviating lipolysis and downregulating lipogenesis through the sirtuin1/adenosine monophosphate-activated protein kinase (Sirt1/AMPK) pathway in lipid-loaded hepatocytes [43].

4. Preventive Effect of Fucoxanthin on NASH Development

NAFLD is a progressive disease that may develop into non-alcoholic steatohepatitis (NASH), fibrosis, cirrhosis, and hepatocellular carcinoma. When it develops into NASH, fibrosis is one of its characteristics and the main cause of mortality [99]. Fucoxanthin not only exhibits a beneficial effect on NAFLD but also has a preventive effect on averting its development into NASH through fibrosis. The regulation of chemokine production suppresses hepatic inflammation and infiltration of immune cells, which is the leading cause of NASH development.
Fucoxanthin is reported to downregulate the hepatic mRNA expression of Tgfβ1, collagen type I alpha 1 chain (Col1α1), and Timp1 in the CDAHFD-fed mouse model [42]. Tgfβ1, transforming growth factor beta 1, is a pro-fibrogenic cytokine that expresses α-smooth muscle actin (αSMA) and promotes extracellular matrix (ECM) production [100]. The suppression of Tgfβ1 inhibits the production of ECM, which prevents fibrosis development. Tgfβ1-induced fibrosis is strongly correlated with matrix metalloproteinase-1 (TIMP-1) expression [101]. Timp1 plays a role in ECM degradation, where the administration of an anti-TIMP-1 antibody ameliorated fibrosis in mice [42].
Fucoxanthin was also reported to downregulate TGFβ1-induced mRNA levels of fibrogenic genes in LX-2 cells. It alleviated the phosphorylation of SMA- and MAD-related protein (SMAD3), which inhibits fibrosis. It exhibited a synergistic effect with SIS3 (an inhibitor of SMAD3) in suppressing fibrogenic gene expression. A similar result has also been reported in hepatic stellate cells. Its anti-fibrogenic activity is further explained through the repression of the NADPH oxidase 4 (NOX4) mRNA levels, which prevented the accumulation of ROS by TGFβ1 [102]. Hence, fucoxanthin exhibits anti-fibrogenic activity that prevents NASH development.

5. Signaling Pathways Altered by Fucoxanthin

Fucoxanthin alters several pathways related to NAFLD, as shown in Figure 5 and Table 2. β3Ad is a metabolic receptor in adipose tissues. The upregulation of β3Ad has been closely related to thermogenesis [103]. A PPAR γ coactivator, PGC-1, is a transcription cofactor that plays a role in regulating cell metabolism [104]. Both PGC-1 and β3Ad stimulate adaptive thermogenesis and mitochondrial biogenesis that favor anti-obesity activity against NAFLD.
The AMPK signaling pathway regulates several metabolic organs, such as the liver, skeletal muscle, pancreas, and adipose tissues. AMPK pathways regulate glucose transport and fatty acid oxidation in skeletal muscle. They upregulate fatty acid oxidation in the liver while decreasing cholesterol and triglyceride synthesis [112]. Several studies have shown that the activation of AMPK results in acetyl-CoA carboxylase suppression, hence blocking fatty acid synthase and decreasing hepatocytic lipid accumulation [43]. The activation of AMPK specific to the liver has been reported to reduce liver steatosis, inflammation, and fibrosis in NAFLD patients. Liver-specific activation of AMPK made mice resistant to weight gain and reduced the overall level of lipid accumulation [113]. Fucoxanthin upregulates AMPK, hence promoting fatty acid oxidation to protect against NAFLD.
Other transcription factors related to lipid metabolism, such as PPARs and sterol regulatory element binding protein (SREBP), are also altered. Similar to AMPK, SREBP regulates the expression of lipogenic enzymes, including fatty acid synthase, acetyl-CoA carboxylase, and 3-hydroxy-3-methylglutaryl-CoA reductase [114]. SREBP-1C is one of the major transcriptional factors involved in de novo lipid synthesis, which affects NAFLD through the nuclear transcription factor farnesoid X receptor (FXR) [41]. Along with stearoyl coenzyme-A desaturase 1 and fatty acid synthase, activation of SREBP-1C increases the rate of fatty acid synthesis. The overexpression of SREBP-1C results in the upregulation of lipogenesis; meanwhile, inactivation of the SREBP-1C gene can reduced triglyceride levels up to 50% in the ob/ob mouse model [115]. Fucoxanthin has been reported to downregulate SREBP-1C; hence, it is believed to reduce lipogenesis and is beneficial for treating NAFLD.
Meanwhile, PPARs have an important role in regulating glucose levels and homeostasis, as well as regulating cell proliferation, differentiation, and inflammation [116]. PPARs are sensors of fatty acids and have tissue-specific expression patterns. PPAR-α is mainly expressed in brown adipose tissues and the liver and regulates lipid metabolism. PPAR-α alteration may lead to hepatic steatosis. Meanwhile, PPAR-β regulates oxidative metabolism (β-oxidation of fatty acids). PPAR-γ is mainly expressed in adipose tissues and macrophages to regulate adipogenesis and storage of fatty acid as triacylglycerol [117]. The activation of PPAR-γ is closely related to obesity, excess nutrients, and the storage of fatty acids as lipids [118]. Fucoxanthin upregulates PPAR-α and PPAR-β while downregulating PPAR-γ [7,9,76]. It functions by upregulating fatty acid β-oxidation and downregulating lipid storage as triacylglycerol in the liver. Fucoxanthin supplementation can alter the PPAR pathway in a favorable manner against NAFLD.
Fucoxanthin downregulated NF-κB, MAPK, and AKT signaling pathways in response to inflammation [86,89]. NF-κB is one of these inflammatory signaling pathways [119]. It plays a role in the homeostasis and expressing the immune response during the cell cycle [120]. Along with FOXP3, NF-κB regulates the secretion of inflammatory cytokines and chemokines. It is a major inducible transcription factor and primarily a cytoplasmic factor expressed in most types of cells [121]. Downregulation of this signaling pathway is related to the alleviation of liver inflammation and improvement in liver histopathology, as shown in rats with type 2 diabetes mellitus (T2DM) and NAFLD treated with liraglutide or hUC-MSCs [122]. Mice with NAFLD induced by a methionine–choline deficient diet (MCDD) showed that both the NF-κB and AKT signaling pathways were downregulated in response to lower levels of inflammation [123].
MAPK is another fundamental inflammation signaling pathway that drives expression of nuclear factor E2-related factor 2 (Nrf2) and NF-κB [124]. MAPK has been found to be responsible for lipid accumulation, inflammation, and ROS production in the HFD mouse model [125]. The suppression of this pathway reduces the inflammatory response in many diseases. The AKT pathway plays a role in cell metabolism, mainly in glucose metabolism. It is also closely associated with cancer and diabetes [126]. Alteration of this pathway is also related to obesity [127]. Fucoxanthin plays a valuable role in downregulating these three major inflammatory signaling pathways.
The anti-oxidant activity of fucoxanthin is closely related to the Nrf2 and AMPK signaling pathway. Fucoxanthin activates the Nrf2 and AMPK signaling pathway to reduce oxidative stress [109,110]. Activation of the AMPK pathway in the liver has been shown to improve the mitochondria’s ability to resist oxidative damage. Meanwhile, the anti-fibrogenic activity of fucoxanthin is mainly through the inhibition of TGF-β1. One of the TGF-β1-related signaling pathways is the SMAD signaling pathway. Fucoxanthin supplementation inhibited TGF-β1 and altered the TGF-β1-dependent SMAD, MAPK signaling pathways, and the phosphatidylinositol 3-kinase (PI3K)/Akt pathway [111]. A recent study using an HPF cell model showed that fucoxantin’s anti-fibrogenic inhibition of TGF-β1 is critically dependent on the SMAD2/3 signaling pathway.

6. Conclusions and Further Potential of Fucoxanthin against NAFLD

The biological activities of fucoxanthin against NAFLD hold promising prospects. However, there are limitations in commercialization due to the high cost of extraction with low yield, low bioavailability, and instability. Various delivery systems for fucoxanthin are currently being developed to increase its bioavailability. Encapsulated fucoxanthin has been proven to have a higher bioavailability in several recent studies [128]. Encapsulation has been attempted using hydroxypropyl-β-cyclodextrin, gum arabic, maltodextrin, gelatin, isolated pea protein, whey protein, zein mixed with caseinate, κ-carrageenan, and poly(d,l-lactic-co-glycolic acid) [129,130,131]. All encapsulation efficiencies were higher than 80%. In addition to encapsulation using a single material, the combination of a protein and polysaccharide is emerging as a major trend for complex delivery systems. Several complex encapsulations have been conducted using arabic/gelatin, whey protein isolate and Ca2+ cross-linked flaxseed gum, zein/chitosan, and nano-encapsulation using gliadin and chondroitin sulfate [132,133,134]. Protein–protein combinations such as lysozyme, protein–lipid combinations such as bovine serum albumin (BSA) and oleic acid, and polysaccharide–lipid combinations such as chitosan–bacuri butter and tucumã oil have also been reported [135,136,137].
Fucoxanthin functions against NAFLD through its thermogenic activity in mitochondrial homeostasis, altering lipid metabolism, its anti-inflammatory and anti-oxidant activities. The thermogenesis activity of fucoxanthin functions against NAFLD via UCP1 activation. Recently, UCP1 has been expected to form a complex with mitochondrial calcium uniporter (MCU) and essential MCU regulator (EMRE), named the thermoporter [40]. While UCP1 is associated with proton leakage, MCU and EMRE are associated with calcium uptake in the mitochondria membrane. As fucoxanthin is known for its thermogenic effect through UCP-1, if UCP-1 indeed forms a complex with EMRE/MCU, mitochondrial calcium uptake regulation may affect UCP-1 and its thermogenic activity. Thus, fucoxanthin’s thermogenic activity could be optimized by finding a compound that synergistically regulates calcium uptake.
Lipid metabolism in the human body is actively altered by changes in diet. De novo lipogenesis plays a major role in NAFLD. It leads to hepatocytic accumulation of triglycerides [73]. Polyunsaturated fatty acid supplementation suppresses lipogenesis gene expression in the liver, including fatty acid synthase, spot14, and stearoyl-CoA desaturase [138]. Meanwhile, a high-carbohydrate diet stimulates lipogenesis in adipose tissues as well as the liver, as indicated by the elevated level of triglycerides [139]. This mechanism can be utilized to find a compound works synergistically with fucoxanthin that can help fight against NAFLD. The conversion step of acetyl-CoA to malonyl-CoA by acetyl-CoA carboxylase is the rate-limiting step in de novo lipogenesis, where malonyl-CoA also plays a role in regulating mitochondrial fat oxidation by inhibiting carnitine palmitoyltransferase I [73]. Soraphen A, a natural polyketide compound isolated from the bacterium Sorangium cellulosum, has been reported to inhibit acetyl-CoA carboxylase and to increase insulin sensitivity in an HFD-fed, insulin-resistant mouse model. Piperidinyl derivative CP-610431, spirocyclic spiropiperidine-derived compound, olumacostat glasaretil, aryl ether-derived analog, piperazine oxadiazole, and 1,4-disubstituted cyclohexane are also other reported acetyl-CoA carboxylase inhibitors [140]. Finding a synergistic inhibitor of acetyl-CoA carboxylase may help fucoxanthin in combating NAFLD.
Further studies on the role of fucoxanthin against NAFLD should also be performed at the level of RNA. Many RNA studies have been conducted regarding NAFLD because of its role in signaling pathways [141]. Non-coding RNA can effectively silence gene expression and alter signaling pathways. A long non-coding RNA highly upregulated in liver cancer (lncRNA HULC) has been found to inhibit the MAPK signaling pathway in an NAFLD mouse model [142]. Locked nucleic acid (LNA) was also found to inhibit lipogenesis and upregulate fatty acid oxidation in db/db mice [143]. A microRNA, miR-291b-3p, affects the AMPK signaling pathway by inhibiting fatty acid synthesis as well as de novo lipogenesis [144]. The microRNA miR-378 alters the AKT signaling pathway and reduces lipogenesis [145]. In another study of 16S rRNA, it was found that fucoxanthin altered high-fat-diet-induced gut microbiota dysbiosis by suppressing the growth of obesityinflammation-related Lachnospiraceae and Erysipelotrichaceae gut bacteria and inducing the growth of Lactobacillus/Lactococcus, Bifidobacterium, and butyrate-producing gut bacteria [146]. A similar study also showed that fucoxanthin altered the Firmicutes/Bacteroidetes ratio and the abundance of S24-7 and Akkermansia, which attenuates obesity in a high fat diet mouse model [147]. Fucoxanthin regulates gut microbiota to treat NAFLD through its anti-obesity activity.

Author Contributions

Conceptualization, C.-H.P.; writing—original draft preparation, J.W.; writing—review and editing, C.-H.P. and D.-G.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Technology Innovation Program (20020268, Establishment of industrialization foundation through development of high value-added health functional food materials and combination formulations using Tisochrysis lutea and Phaeodactylum tricornutum) funded by the Ministry of Trade, Industry & Energy (MOTIE, Korea) and was funded by the Korea Institute of Marine Science & Technology Promotion (KIMST) funded by the Ministry of Oceans and Fisheries (grant No. 20220134 and grant No. 20220027).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created.

Acknowledgments

All figures in this paper were created using BioRender.com (accessed on 23 November 2022). Some templates were adjusted accordingly for visualization purposes and cited in the references.

Conflicts of Interest

Microalgae Ask US Co., Ltd., one of the affiliated institutions of the authors of this paper, is an institution that conducts fucoxanthin-related research with funding from MOTIE and KIMST. Aside from this, the authors have no conflicts of interest.

References

  1. Cheemerla, S.; Balakrishnan, M. Global Epidemiology of Chronic Liver Disease. Clin. Liver Dis. 2021, 17, 365–370. [Google Scholar] [CrossRef] [PubMed]
  2. Collaborators, G.B.D.C. The global, regional, and national burden of cirrhosis by cause in 195 countries and territories, 1990–2017: A systematic analysis for the Global Burden of Disease Study 2017. Lancet Gastroenterol. Hepatol. 2020, 5, 245–266. [Google Scholar]
  3. Moon, A.M.; Singal, A.G.; Tapper, E.B. Contemporary Epidemiology of Chronic Liver Disease and Cirrhosis. Clin. Gastroenterol. Hepatol. 2020, 18, 2650–2666. [Google Scholar] [CrossRef] [PubMed]
  4. Ramanathan, R.; Ali, A.H.; Ibdah, J.A. Mitochondrial Dysfunction Plays Central Role in Nonalcoholic Fatty Liver Disease. Int. J. Mol. Sci. 2022, 23, 7280. [Google Scholar] [CrossRef] [PubMed]
  5. D’Orazio, N.; Gemello, E.; Gammone, M.A.; De Girolamo, M.; Ficoneri, C.; Riccioni, G. Fucoxantin: A treasure from the sea. Mar. Drugs 2012, 10, 604–616. [Google Scholar] [CrossRef]
  6. Gammone, M.A.; D’Orazio, N. Anti-obesity activity of the marine carotenoid fucoxanthin. Mar. Drugs 2015, 13, 2196–2214. [Google Scholar] [CrossRef]
  7. Kang, S.I.; Shin, H.S.; Kim, H.M.; Yoon, S.A.; Kang, S.W.; Kim, J.H.; Ko, H.C.; Kim, S.J. Petalonia binghamiae extract and its constituent fucoxanthin ameliorate high-fat diet-induced obesity by activating AMP-activated protein kinase. J. Agric. Food Chem. 2012, 60, 3389–3395. [Google Scholar] [CrossRef]
  8. Koo, S.Y.; Hwang, J.H.; Yang, S.H.; Um, J.I.; Hong, K.W.; Kang, K.; Pan, C.H.; Hwang, K.T.; Kim, S.M. Anti-Obesity Effect of Standardized Extract of Microalga Phaeodactylum tricornutum Containing Fucoxanthin. Mar. Drugs 2019, 17, 311. [Google Scholar] [CrossRef]
  9. Woo, M.N.; Jeon, S.M.; Kim, H.J.; Lee, M.K.; Shin, S.K.; Shin, Y.C.; Park, Y.B.; Choi, M.S. Fucoxanthin supplementation improves plasma and hepatic lipid metabolism and blood glucose concentration in high-fat fed C57BL/6N mice. Chem. -Biol. Interact. 2010, 186, 316–322. [Google Scholar] [CrossRef] [PubMed]
  10. Li, H.; Ji, H.; Shi, C.; Gao, Y.; Zhang, Y.; Xu, X.; Ding, H.; Tang, L.; Xing, Y. Distribution of heavy metals and metalloids in bulk and particle size fractions of soils from coal-mine brownfield and implications on human health. Chemosphere 2017, 172, 505–515. [Google Scholar] [CrossRef]
  11. Mikami, K.; Hosokawa, M. Biosynthetic pathway and health benefits of fucoxanthin, an algae-specific xanthophyll in brown seaweeds. Int. J. Mol. Sci. 2013, 14, 13763–13781. [Google Scholar] [CrossRef]
  12. Lourenço-Lopes, C.; Fraga-Corral, M.; Jimenez-Lopez, C.; Carpena, M.; Pereira, A.; García-Oliveira, P.; Prieto, M.; Simal-Gandara, J. Biological action mechanisms of fucoxanthin extracted from algae for application in food and cosmetic industries. Trends Food Sci. Technol. 2021, 117, 163–181. [Google Scholar] [CrossRef]
  13. Murase, W.; Kamakura, Y.; Kawakami, S.; Yasuda, A.; Wagatsuma, M.; Kubota, A.; Kojima, H.; Ohta, T.; Takahashi, M.; Mutoh, M.; et al. Fucoxanthin Prevents Pancreatic Tumorigenesis in C57BL/6J Mice That Received Allogenic and Orthotopic Transplants of Cancer Cells. Int. J. Mol. Sci. 2021, 22, 13620. [Google Scholar] [CrossRef] [PubMed]
  14. Terasaki, M.; Kubota, A.; Kojima, H.; Maeda, H.; Miyashita, K.; Kawagoe, C.; Mutoh, M.; Tanaka, T. Fucoxanthin and Colorectal Cancer Prevention. Cancers 2021, 13, 2379. [Google Scholar] [CrossRef]
  15. Hu, T.; Liu, D.; Chen, Y.; Wu, J.; Wang, S. Antioxidant activity of sulfated polysaccharide fractions extracted from Undaria pinnitafida in vitro. Int. J. Biol. Macromol. 2010, 46, 193–198. [Google Scholar] [CrossRef]
  16. Peng, J.; Yuan, J.P.; Wu, C.F.; Wang, J.H. Fucoxanthin, a marine carotenoid present in brown seaweeds and diatoms: Metabolism and bioactivities relevant to human health. Mar. Drugs 2011, 9, 1806–1828. [Google Scholar] [CrossRef] [PubMed]
  17. Miyashita, K.; Nishikawa, S.; Beppu, F.; Tsukui, T.; Abe, M.; Hosokawa, M. The allenic carotenoid fucoxanthin, a novel marine nutraceutical from brown seaweeds. J. Sci. Food Agric. 2011, 91, 1166–1174. [Google Scholar] [CrossRef]
  18. Sachindra, N.M.; Sato, E.; Maeda, H.; Hosokawa, M.; Niwano, Y.; Kohno, M.; Miyashita, K. Radical scavenging and singlet oxygen quenching activity of marine carotenoid fucoxanthin and its metabolites. J. Agric. Food Chem. 2007, 55, 8516–8522. [Google Scholar] [CrossRef]
  19. Bae, M.; Kim, M.B.; Park, Y.K.; Lee, J.Y. Health benefits of fucoxanthin in the prevention of chronic diseases. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2020, 1865, 158618. [Google Scholar] [CrossRef] [PubMed]
  20. Kajikawa, T.; Okumura, S.; Iwashita, T.; Kosumi, D.; Hashimoto, H.; Katsumura, S. Stereocontrolled total synthesis of fucoxanthin and its polyene chain-modified derivative. Org. Lett. 2012, 14, 808–811. [Google Scholar] [CrossRef]
  21. Stahl, W.; Sies, H. Antioxidant activity of carotenoids. Mol. Asp. Med. 2003, 24, 345–351. [Google Scholar] [CrossRef]
  22. Conn, P.F.; Schalch, W.; Truscott, T.G. The singlet oxygen and carotenoid interaction. J. Photochem. Photobiol. B Biol. 1991, 11, 41–47. [Google Scholar] [CrossRef]
  23. Rajauria, G.; Foley, B.; Abu-Ghannam, N. Characterization of dietary fucoxanthin from Himanthalia elongata brown seaweed. Food Res. Int. 2017, 99 Pt 3, 995–1001. [Google Scholar] [CrossRef]
  24. Miller, N.J.; Sampson, J.; Candeias, L.P.; Bramley, P.M.; Rice-Evans, C.A. Antioxidant activities of carotenes and xanthophylls. FEBS Lett. 1996, 384, 240–242. [Google Scholar] [CrossRef]
  25. Maeda, H. Anti-obesity and anti-diabetic activities of algae. In Functional Ingredients from Algae for Foods and Nutraceuticals; Elsevier: Amsterdam, The Netherlands, 2013; pp. 453–472. [Google Scholar]
  26. Soriano, E.; Fernandez, I. Allenes and computational chemistry: From bonding situations to reaction mechanisms. Chem. Soc. Rev. 2014, 43, 3041–3105. [Google Scholar] [CrossRef] [PubMed]
  27. Sangeetha, R.K.; Bhaskar, N.; Divakar, S.; Baskaran, V. Bioavailability and metabolism of fucoxanthin in rats: Structural characterization of metabolites by LC-MS (APCI). Mol. Cell Biochem. 2010, 333, 299–310. [Google Scholar] [CrossRef]
  28. Asai, A.; Sugawara, T.; Ono, H.; Nagao, A. Biotransformation of fucoxanthinol into amarouciaxanthin A in mice and HepG2 cells: Formation and cytotoxicity of fucoxanthin metabolites. Drug Metab. Dispos. 2004, 32, 205–211. [Google Scholar] [CrossRef] [PubMed]
  29. Das, S.K.; Hashimoto, T.; Kanazawa, K. Growth inhibition of human hepatic carcinoma HepG2 cells by fucoxanthin is associated with down-regulation of cyclin D. Biochim. Biophys Acta 2008, 1780, 743–749. [Google Scholar] [CrossRef]
  30. Hashimoto, T.; Ozaki, Y.; Taminato, M.; Das, S.K.; Mizuno, M.; Yoshimura, K.; Maoka, T.; Kanazawa, K. The distribution and accumulation of fucoxanthin and its metabolites after oral administration in mice. Br. J. Nutr. 2009, 102, 242–248. [Google Scholar] [CrossRef] [PubMed]
  31. Beppu, F.; Niwano, Y.; Tsukui, T.; Hosokawa, M.; Miyashita, K. Single and repeated oral dose toxicity study of fucoxanthin (FX), a marine carotenoid, in mice. J. Toxicol. Sci. 2009, 34, 501–510. [Google Scholar] [CrossRef]
  32. Iio, K.; Okada, Y.; Ishikura, M. Single and 13-week oral toxicity study of fucoxanthin oil from microalgae in rats. J. Food Hyg. Soc. Jpn. 2011, 52, 183–189. [Google Scholar] [CrossRef] [PubMed]
  33. Spagolla Napoleao Tavares, R.; Maria-Engler, S.S.; Colepicolo, P.; Debonsi, H.M.; Schafer-Korting, M.; Marx, U.; Gaspar, L.R.; Zoschke, C. Skin Irritation Testing beyond Tissue Viability: Fucoxanthin Effects on Inflammation, Homeostasis, and Metabolism. Pharmaceutics 2020, 12, 136. [Google Scholar] [CrossRef] [PubMed]
  34. Fabbrini, E.; Sullivan, S.; Klein, S. Obesity and nonalcoholic fatty liver disease: Biochemical, metabolic, and clinical implications. Hepatology 2010, 51, 679–689. [Google Scholar] [CrossRef] [PubMed]
  35. Marchesini, G.; Bugianesi, E.; Forlani, G.; Cerrelli, F.; Lenzi, M.; Manini, R.; Natale, S.; Vanni, E.; Villanova, N.; Melchionda, N.; et al. Nonalcoholic fatty liver, steatohepatitis, and the metabolic syndrome. Hepatology 2003, 37, 917–923. [Google Scholar] [CrossRef]
  36. Oliveros, E.; Somers, V.K.; Sochor, O.; Goel, K.; Lopez-Jimenez, F. The concept of normal weight obesity. Prog. Cardiovasc. Dis. 2014, 56, 426–433. [Google Scholar] [CrossRef] [PubMed]
  37. Smith, U. Abdominal obesity: A marker of ectopic fat accumulation. J. Clin. Investig. 2015, 125, 1790–1792. [Google Scholar] [CrossRef]
  38. Polyzos, S.A.; Kountouras, J.; Mantzoros, C.S. Obesity and nonalcoholic fatty liver disease: From pathophysiology to therapeutics. Metabolism 2019, 92, 82–97. [Google Scholar] [CrossRef]
  39. Li, L.; Liu, D.W.; Yan, H.Y.; Wang, Z.Y.; Zhao, S.H.; Wang, B. Obesity is an independent risk factor for non-alcoholic fatty liver disease: Evidence from a meta-analysis of 21 cohort studies. Obes. Rev. 2016, 17, 510–519. [Google Scholar] [CrossRef]
  40. Xue, K.; Wu, D.; Wang, Y.; Zhao, Y.; Shen, H.; Yao, J.; Huang, X.; Li, X.; Zhou, Z.; Wang, Z.; et al. The mitochondrial calcium uniporter engages UCP1 to form a thermoporter that promotes thermogenesis. Cell Metab. 2022, 34, 1325–1341.e6. [Google Scholar] [CrossRef]
  41. Friedman, S.L.; Neuschwander-Tetri, B.A.; Rinella, M.; Sanyal, A.J. Mechanisms of NAFLD development and therapeutic strategies. Nat. Med. 2018, 24, 908–922. [Google Scholar] [CrossRef]
  42. Takatani, N.; Kono, Y.; Beppu, F.; Okamatsu-Ogura, Y.; Yamano, Y.; Miyashita, K.; Hosokawa, M. Fucoxanthin inhibits hepatic oxidative stress, inflammation, and fibrosis in diet-induced nonalcoholic steatohepatitis model mice. Biochem. Biophys. Res. Commun. 2020, 528, 305–310. [Google Scholar] [CrossRef]
  43. Chang, Y.H.; Chen, Y.L.; Huang, W.C.; Liou, C.J. Fucoxanthin attenuates fatty acid-induced lipid accumulation in FL83B hepatocytes through regulated Sirt1/AMPK signaling pathway. Biochem. Biophys Res. Commun. 2018, 495, 197–203. [Google Scholar] [CrossRef]
  44. Gusdon, A.M.; Song, K.X.; Qu, S. Nonalcoholic Fatty liver disease: Pathogenesis and therapeutics from a mitochondria-centric perspective. Oxidative Med. Cell. Longev. 2014, 2014, 637027. [Google Scholar] [CrossRef] [PubMed]
  45. Li, Z.; Li, Y.; Zhang, H.X.; Guo, J.R.; Lam, C.W.K.; Wang, C.Y.; Zhang, W. Mitochondria-Mediated Pathogenesis and Therapeutics for Non-Alcoholic Fatty Liver Disease. Mol. Nutr. Food Res. 2019, 63, e1900043. [Google Scholar] [CrossRef]
  46. Mitchell, P. Vectorial chemistry and the molecular mechanics of chemiosmotic coupling: Power transmission by proticity. Biochem. Soc. Trans. 1976, 4, 399–430. [Google Scholar] [CrossRef]
  47. Smith, R.E.; Roberts, J.C.; Hittelman, K.J. Nonphosphorylating respiration of mitochondria from brown adipose tissue of rats. Science 1966, 154, 653–654. [Google Scholar] [CrossRef] [PubMed]
  48. Divakaruni, A.S.; Brand, M.D. The regulation and physiology of mitochondrial proton leak. Physiology 2011, 26, 192–205. [Google Scholar] [CrossRef]
  49. Jastroch, M.; Divakaruni, A.S.; Mookerjee, S.; Treberg, J.R.; Brand, M.D. Mitochondrial proton and electron leaks. Essays Biochem. 2010, 47, 53–67. [Google Scholar]
  50. Huang, E. Different Types of Membrane Transport. 2022. Available online: https://app.biorender.com/biorender-templates/t-6329c98fbae8d5e25bed4fed-different-types-of-membrane-transport (accessed on 23 November 2022).
  51. Maeda, H.; Hosokawa, M.; Sashima, T.; Funayama, K.; Miyashita, K. Fucoxanthin from edible seaweed, Undaria pinnatifida, shows antiobesity effect through UCP1 expression in white adipose tissues. Biochem. Biophys. Res. Commun. 2005, 332, 392–397. [Google Scholar] [CrossRef] [PubMed]
  52. Cannon, B.; Nedergaard, J. Brown adipose tissue: Function and physiological significance. Physiol. Rev. 2004, 84, 277–359. [Google Scholar] [CrossRef]
  53. Smorlesi, A.; Frontini, A.; Giordano, A.; Cinti, S. The adipose organ: White-brown adipocyte plasticity and metabolic inflammation. Obes. Rev. 2012, 13 (Suppl. S2), 83–96. [Google Scholar] [CrossRef] [PubMed]
  54. Fenzl, A.; Kiefer, F.W. Brown adipose tissue and thermogenesis. Horm. Mol. Biol. Clin. Investig. 2014, 19, 25–37. [Google Scholar] [CrossRef]
  55. Symonds, M.E. Brown adipose tissue growth and development. Scientifica 2013, 2013, 305763. [Google Scholar] [CrossRef]
  56. Hosokawa, M.; Miyashita, T.; Nishikawa, S.; Emi, S.; Tsukui, T.; Beppu, F.; Okada, T.; Miyashita, K. Fucoxanthin regulates adipocytokine mRNA expression in white adipose tissue of diabetic/obese KK-Ay mice. Arch. Biochem. Biophys. 2010, 504, 17–25. [Google Scholar] [CrossRef]
  57. Maeda, H.; Hosokawa, M.; Sashima, T.; Murakami-Funayama, K.; Miyashita, K. Anti-obesity and anti-diabetic effects of fucoxanthin on diet-induced obesity conditions in a murine model. Mol. Med.Rep. 2009, 2, 897–902. [Google Scholar] [CrossRef]
  58. Enerback, S. Human brown adipose tissue. Cell Metab. 2010, 11, 248–252. [Google Scholar] [CrossRef]
  59. Seale, P. Transcriptional control of brown adipocyte development and thermogenesis. Int. J. Obes. 2010, 34 (Suppl. S1), S17–S22. [Google Scholar] [CrossRef]
  60. Miyashita, K. Anti-obesity therapy by food component: Unique activity of marine carotenoid, fucoxanthin. Obes. Control. 2014, 1, 4. [Google Scholar] [CrossRef]
  61. Maeda, H.; Hosokawa, M.; Sashima, T.; Funayama, K.; Miyashita, K. Effect of medium-chain triacylglycerols on anti-obesity effect of fucoxanthin. J. Oleo. Sci. 2007, 56, 615–621. [Google Scholar] [CrossRef]
  62. Ahn, J.; Ha, T.Y.; Ahn, J.; Jung, C.H.; Seo, H.D.; Kim, M.J.; Kim, Y.S.; Jang, Y.J. Undaria pinnatifida extract feeding increases exercise endurance and skeletal muscle mass by promoting oxidative muscle remodeling in mice. FASEB J. 2020, 34, 8068–8081. [Google Scholar] [CrossRef] [PubMed]
  63. Jia, J.J.; Tian, Y.B.; Cao, Z.H.; Tao, L.L.; Zhang, X.; Gao, S.Z.; Ge, C.R.; Lin, Q.Y.; Jois, M. The polymorphisms of UCP1 genes associated with fat metabolism, obesity and diabetes. Mol. Biol. Rep. 2010, 37, 1513–1522. [Google Scholar] [CrossRef] [PubMed]
  64. Kersten, S. Mechanisms of nutritional and hormonal regulation of lipogenesis. EMBO Rep. 2001, 2, 282–286. [Google Scholar] [CrossRef] [PubMed]
  65. Deprince, A.; Haas, J.T.; Staels, B. Dysregulated lipid metabolism links NAFLD to cardiovascular disease. Mol. Metab. 2020, 42, 101092. [Google Scholar] [CrossRef]
  66. Hodson, L.; Gunn, P.J. The regulation of hepatic fatty acid synthesis and partitioning: The effect of nutritional state. Nat. Rev. Endocrinol. 2019, 15, 689–700. [Google Scholar] [CrossRef]
  67. Nassir, F.; Rector, R.S.; Hammoud, G.M.; Ibdah, J.A. Pathogenesis and Prevention of Hepatic Steatosis. Gastroenterol. Hepatol. 2015, 11, 167–175. [Google Scholar]
  68. Muradian, K.; Vaiserman, A.; Min, K.J.; Fraifeld, V.E. Fucoxanthin and lipid metabolism: A minireview. Nutr. Metab. Cardiovasc. Dis. 2015, 25, 891–897. [Google Scholar] [CrossRef]
  69. Woo, M.N.; Jeon, S.M.; Shin, Y.C.; Lee, M.K.; Kang, M.A.; Choi, M.S. Anti-obese property of fucoxanthin is partly mediated by altering lipid-regulating enzymes and uncoupling proteins of visceral adipose tissue in mice. Mol. Nutr. Food Res 2009, 53, 1603–1611. [Google Scholar] [CrossRef]
  70. Takenaka, A.; Nakamura, S.; Mitsunaga, F.; Inoue-Murayama, M.; Udono, T.; Suryobroto, B. Human-specific SNP in obesity genes, adrenergic receptor beta2 (ADRB2), Beta3 (ADRB3), and PPAR gamma2 (PPARG), during primate evolution. PLoS ONE 2012, 7, e43461. [Google Scholar] [CrossRef]
  71. Fang, H.; Judd, R.L. Adiponectin Regulation and Function. Compr. Physiol. 2018, 8, 1031–1063. [Google Scholar]
  72. Kadowaki, T.; Yamauchi, T. Adiponectin and adiponectin receptors. Endocr. Rev. 2005, 26, 439–451. [Google Scholar] [CrossRef] [PubMed]
  73. Alkhouri, N.; Lawitz, E.; Noureddin, M.; DeFronzo, R.; Shulman, G.I. GS-0976 (Firsocostat): An investigational liver-directed acetyl-CoA carboxylase (ACC) inhibitor for the treatment of non-alcoholic steatohepatitis (NASH). Expert Opin. Investig. Drugs 2020, 29, 135–141. [Google Scholar] [CrossRef] [PubMed]
  74. Kang, S.I.; Kim, M.H.; Shin, H.S.; Kim, H.M.; Hong, Y.S.; Park, J.G.; Ko, H.C.; Lee, N.H.; Chung, W.S.; Kim, S.J. A water-soluble extract of Petalonia binghamiae inhibits the expression of adipogenic regulators in 3T3-L1 preadipocytes and reduces adiposity and weight gain in rats fed a high-fat diet. J. Nutr. Biochem. 2010, 21, 1251–1257. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, Y.; Xu, W.; Huang, X.; Zhao, Y.; Ren, Q.; Hong, Z.; Huang, M.; Xing, X. Fucoxanthin ameliorates hyperglycemia, hyperlipidemia and insulin resistance in diabetic mice partially through IRS-1/PI3K/Akt and AMPK pathways. J. Funct. Foods 2018, 48, 515–524. [Google Scholar] [CrossRef]
  76. Park, H.J.; Lee, M.K.; Park, Y.B.; Shin, Y.C.; Choi, M.S. Beneficial effects of Undaria pinnatifida ethanol extract on diet-induced-insulin resistance in C57BL/6J mice. Food Chem. Toxicol. 2011, 49, 727–733. [Google Scholar] [CrossRef]
  77. Shih, P.H.; Shiue, S.J.; Chen, C.N.; Cheng, S.W.; Lin, H.Y.; Wu, L.W.; Wu, M.S. Fucoidan and Fucoxanthin Attenuate Hepatic Steatosis and Inflammation of NAFLD through Modulation of Leptin/Adiponectin Axis. Mar. Drugs 2021, 19, 148. [Google Scholar] [CrossRef] [PubMed]
  78. Kim, M.B.; Bae, M.; Lee, Y.; Kang, H.; Hu, S.; Pham, T.X.; Park, Y.K.; Lee, J.Y. Consumption of Low Dose Fucoxanthin Does Not Prevent Hepatic and Adipose Inflammation and Fibrosis in Mouse Models of Diet-Induced Obesity. Nutrients 2022, 14, 2280. [Google Scholar] [CrossRef]
  79. Tan, C.P.; Hou, Y.H. First evidence for the anti-inflammatory activity of fucoxanthin in high-fat-diet-induced obesity in mice and the antioxidant functions in PC12 cells. Inflammation 2014, 37, 443–450. [Google Scholar] [CrossRef] [PubMed]
  80. Katsarou, A.; Moustakas, I.I.; Pyrina, I.; Lembessis, P.; Koutsilieris, M.; Chatzigeorgiou, A. Metabolic inflammation as an instigator of fibrosis during non-alcoholic fatty liver disease. World J. Gastroenterol. 2020, 26, 1993–2011. [Google Scholar] [CrossRef] [PubMed]
  81. Świderska, M.; Maciejczyk, M.; Zalewska, A.; Pogorzelska, J.; Flisiak, R.; Chabowski, A. Oxidative stress biomarkers in the serum and plasma of patients with non-alcoholic fatty liver disease (NAFLD). Can plasma AGE be a marker of NAFLD? Oxidative stress biomarkers in NAFLD patients. Free Radic. Res. 2019, 53, 841–850. [Google Scholar] [CrossRef]
  82. Kitade, H.; Chen, G.; Ni, Y.; Ota, T. Nonalcoholic Fatty Liver Disease and Insulin Resistance: New Insights and Potential New Treatments. Nutrients 2017, 9, 387. [Google Scholar] [CrossRef]
  83. Hajighasem, A.; Farzanegi, P.; Mazaheri, Z. Effects of combined therapy with resveratrol, continuous and interval exercises on apoptosis, oxidative stress, and inflammatory biomarkers in the liver of old rats with non-alcoholic fatty liver disease. Arch. Physiol. Biochem. 2019, 125, 142–149. [Google Scholar] [CrossRef]
  84. El-Din, S.H.; Sabra, A.N.; Hammam, O.A.; Ebeid, F.A.; El-Lakkany, N.M. Pharmacological and antioxidant actions of garlic and.or onion in non-alcoholic fatty liver disease (NAFLD) in rats. J. Egypt. Soc. Parasitol. 2014, 44, 295–308. [Google Scholar]
  85. Clark, R.B. The role of PPARs in inflammation and immunity. J. Leukoc. Biol. 2002, 71, 388–400. [Google Scholar] [CrossRef]
  86. Kim, K.N.; Heo, S.J.; Yoon, W.J.; Kang, S.M.; Ahn, G.; Yi, T.H.; Jeon, Y.J. Fucoxanthin inhibits the inflammatory response by suppressing the activation of NF-kappaB and MAPKs in lipopolysaccharide-induced RAW 264.7 macrophages. Eur. J. Pharm. 2010, 649, 369–375. [Google Scholar] [CrossRef] [PubMed]
  87. Li, X.; Huang, R.; Liu, K.; Li, M.; Luo, H.; Cui, L.; Huang, L.; Luo, L. Fucoxanthin attenuates LPS-induced acute lung injury via inhibition of the TLR4/MyD88 signaling axis. Aging 2020, 13, 2655–2667. [Google Scholar] [CrossRef]
  88. Su, J.; Guo, K.; Huang, M.; Liu, Y.; Zhang, J.; Sun, L.; Li, D.; Pang, K.L.; Wang, G.; Chen, L.; et al. Fucoxanthin, a Marine Xanthophyll Isolated From Conticribra weissflogii ND-8: Preventive Anti-Inflammatory Effect in a Mouse Model of Sepsis. Front. Pharmacol. 2019, 10, 906. [Google Scholar] [CrossRef] [PubMed]
  89. Choi, J.H.; Kim, N.H.; Kim, S.J.; Lee, H.J.; Kim, S. Fucoxanthin Inhibits the Inflammation Response in Paw Edema Model through Suppressing MAPKs, Akt, and NFkappaB. J. Biochem. Mol. Toxicol. 2016, 30, 111–119. [Google Scholar] [CrossRef]
  90. Wu, S.J.; Liou, C.J.; Chen, Y.L.; Cheng, S.C.; Huang, W.C. Fucoxanthin Ameliorates Oxidative Stress and Airway Inflammation in Tracheal Epithelial Cells and Asthmatic Mice. Cells 2021, 10, 1311. [Google Scholar] [CrossRef] [PubMed]
  91. Bulua, A.C.; Simon, A.; Maddipati, R.; Pelletier, M.; Park, H.; Kim, K.Y.; Sack, M.N.; Kastner, D.L.; Siegel, R.M. Mitochondrial reactive oxygen species promote production of proinflammatory cytokines and are elevated in TNFR1-associated periodic syndrome (TRAPS). J. Exp. Med. 2011, 208, 519–533. [Google Scholar] [CrossRef]
  92. Cruz, C.M.; Rinna, A.; Forman, H.J.; Ventura, A.L.; Persechini, P.M.; Ojcius, D.M. ATP activates a reactive oxygen species-dependent oxidative stress response and secretion of proinflammatory cytokines in macrophages. J. Biol. Chem. 2007, 282, 2871–2879. [Google Scholar] [CrossRef] [PubMed]
  93. Brieger, K.; Schiavone, S.; Miller, F.J., Jr.; Krause, K.H. Reactive oxygen species: From health to disease. Swiss Med. Wkly. 2012, 142, w13659. [Google Scholar] [CrossRef]
  94. Farzanegi, P.; Dana, A.; Ebrahimpoor, Z.; Asadi, M.; Azarbayjani, M.A. Mechanisms of beneficial effects of exercise training on non-alcoholic fatty liver disease (NAFLD): Roles of oxidative stress and inflammation. Eur. J. Sport Sci. 2019, 19, 994–1003. [Google Scholar] [CrossRef] [PubMed]
  95. Polimeni, L.; Del Ben, M.; Baratta, F.; Perri, L.; Albanese, F.; Pastori, D.; Violi, F.; Angelico, F. Oxidative stress: New insights on the association of non-alcoholic fatty liver disease and atherosclerosis. World J. Hepatol. 2015, 7, 1325–1336. [Google Scholar] [CrossRef] [PubMed]
  96. Maeda, H.; Fukuda, S.; Izumi, H.; Saga, N. Anti-Oxidant and Fucoxanthin Contents of Brown Alga Ishimozuku (Sphaerotrichia divaricata) from the West Coast of Aomori, Japan. Mar. Drugs 2018, 16, 255. [Google Scholar] [CrossRef]
  97. Seo, M.J.; Seo, Y.J.; Pan, C.H.; Lee, O.H.; Kim, K.J.; Lee, B.Y. Fucoxanthin Suppresses Lipid Accumulation and ROS Production During Differentiation in 3T3-L1 Adipocytes. Phytother. Res. 2016, 30, 1802–1808. [Google Scholar] [CrossRef]
  98. Zhao, Y.-Q.; Zhang, L.; Zhao, G.-X.; Chen, Y.; Sun, K.-L.; Wang, B. Fucoxanthin attenuates doxorubicin-induced cardiotoxicity via anti-oxidant and anti-apoptotic mechanisms associated with p38, JNK and p53 pathways. J. Funct. Foods 2019, 62, 103542. [Google Scholar] [CrossRef]
  99. Schwabe, R.F.; Tabas, I.; Pajvani, U.B. Mechanisms of Fibrosis Development in Nonalcoholic Steatohepatitis. Gastroenterology 2020, 158, 1913–1928. [Google Scholar] [CrossRef] [PubMed]
  100. Bae, M.; Park, Y.K.; Lee, J.Y. Food components with antifibrotic activity and implications in prevention of liver disease. J. Nutr. Biochem. 2018, 55, 1–11. [Google Scholar] [CrossRef] [PubMed]
  101. Kokeny, G.; Nemeth, A.; Kopp, J.B.; Chen, W.; Oler, A.J.; Manzeger, A.; Rosivall, L.; Mozes, M.M. Susceptibility to kidney fibrosis in mice is associated with early growth response-2 protein and tissue inhibitor of metalloproteinase-1 expression. Kidney Int. 2022, 102, 337–354. [Google Scholar] [CrossRef]
  102. Kim, M.B.; Bae, M.; Hu, S.; Kang, H.; Park, Y.K.; Lee, J.Y. Fucoxanthin exerts anti-fibrogenic effects in hepatic stellate cells. Biochem. Biophys. Res. Commun. 2019, 513, 657–662. [Google Scholar] [CrossRef]
  103. Michel, L.Y.M.; Farah, C.; Balligand, J.L. The Beta3 Adrenergic Receptor in Healthy and Pathological Cardiovascular Tissues. Cells 2020, 9, 2584. [Google Scholar] [CrossRef] [PubMed]
  104. Liang, H.; Ward, W.F. PGC-1alpha: A key regulator of energy metabolism. Adv. Physiol. Educ. 2006, 30, 145–151. [Google Scholar] [CrossRef] [PubMed]
  105. Nishikawa, S.; Hosokawa, M.; Miyashita, K. Fucoxanthin promotes translocation and induction of glucose transporter 4 in skeletal muscles of diabetic/obese KK-A(y) mice. Phytomedicine 2012, 19, 389–394. [Google Scholar] [CrossRef] [PubMed]
  106. Kang, S.I.; Ko, H.C.; Shin, H.S.; Kim, H.M.; Hong, Y.S.; Lee, N.H.; Kim, S.J. Fucoxanthin exerts differing effects on 3T3-L1 cells according to differentiation stage and inhibits glucose uptake in mature adipocytes. Biochem. Biophys. Res. Commun. 2011, 409, 769–774. [Google Scholar] [CrossRef]
  107. Yim, M.J.; Hosokawa, M.; Mizushina, Y.; Yoshida, H.; Saito, Y.; Miyashita, K. Suppressive effects of Amarouciaxanthin A on 3T3-L1 adipocyte differentiation through down-regulation of PPARgamma and C/EBPalpha mRNA expression. J. Agric. Food Chem. 2011, 59, 1646–1652. [Google Scholar] [CrossRef]
  108. Zheng, J.; Tian, X.; Zhang, W.; Zheng, P.; Huang, F.; Ding, G.; Yang, Z. Protective Effects of Fucoxanthin against Alcoholic Liver Injury by Activation of Nrf2-Mediated Antioxidant Defense and Inhibition of TLR4-Mediated Inflammation. Mar. Drugs 2019, 17, 552. [Google Scholar] [CrossRef]
  109. Zheng, D.; Chen, L.; Li, G.; Jin, L.; Wei, Q.; Liu, Z.; Yang, G.; Li, Y.; Xie, X. Fucoxanthin ameliorated myocardial fibrosis in STZ-induced diabetic rats and cell hypertrophy in HG-induced H9c2 cells by alleviating oxidative stress and restoring mitophagy. Food Funct. 2022, 13, 9559–9575. [Google Scholar] [CrossRef] [PubMed]
  110. Jang, E.J.; Kim, S.C.; Lee, J.H.; Lee, J.R.; Kim, I.K.; Baek, S.Y.; Kim, Y.W. Fucoxanthin, the constituent of Laminaria japonica, triggers AMPK-mediated cytoprotection and autophagy in hepatocytes under oxidative stress. BMC Complement. Altern. Med. 2018, 18, 97. [Google Scholar] [CrossRef]
  111. Ma, S.Y.; Park, W.S.; Lee, D.S.; Choi, G.; Yim, M.J.; Lee, J.M.; Jung, W.K.; Park, S.G.; Seo, S.K.; Park, S.J.; et al. Fucoxanthin inhibits profibrotic protein expression in vitro and attenuates bleomycin-induced lung fibrosis in vivo. Eur. J. Pharm. 2017, 811, 199–207. [Google Scholar] [CrossRef] [PubMed]
  112. Gruzman, A.; Babai, G.; Sasson, S. Adenosine Monophosphate-Activated Protein Kinase (AMPK) as a New Target for Antidiabetic Drugs: A Review on Metabolic, Pharmacological and Chemical Considerations. Rev. Diabet. Stud. 2009, 6, 13–36. [Google Scholar] [CrossRef]
  113. Garcia, D.; Hellberg, K.; Chaix, A.; Wallace, M.; Herzig, S.; Badur, M.G.; Lin, T.; Shokhirev, M.N.; Pinto, A.F.M.; Ross, D.S.; et al. Genetic Liver-Specific AMPK Activation Protects against Diet-Induced Obesity and NAFLD. Cell Rep. 2019, 26, 192–208.e6. [Google Scholar] [CrossRef]
  114. Goldstein, J.L.; Rawson, R.B.; Brown, M.S. Mutant mammalian cells as tools to delineate the sterol regulatory element-binding protein pathway for feedback regulation of lipid synthesis. Arch. Biochem. Biophys. 2002, 397, 139–148. [Google Scholar] [CrossRef]
  115. Kim, Y.R.; Lee, E.J.; Shin, K.O.; Kim, M.H.; Pewzner-Jung, Y.; Lee, Y.M.; Park, J.W.; Futerman, A.H.; Park, W.J. Hepatic triglyceride accumulation via endoplasmic reticulum stress-induced SREBP-1 activation is regulated by ceramide synthases. Exp. Mol. Med. 2019, 51, 1–16. [Google Scholar] [CrossRef]
  116. Peters, J.M.; Shah, Y.M.; Gonzalez, F.J. The role of peroxisome proliferator-activated receptors in carcinogenesis and chemoprevention. Nat. Rev. Cancer 2012, 12, 181–195. [Google Scholar] [CrossRef]
  117. Wang, Y.; Nakajima, T.; Gonzalez, F.J.; Tanaka, N. PPARs as Metabolic Regulators in the Liver: Lessons from Liver-Specific PPAR-Null Mice. Int. J. Mol. Sci. 2020, 21, 2061. [Google Scholar] [CrossRef] [PubMed]
  118. Matsusue, K.; Kusakabe, T.; Noguchi, T.; Takiguchi, S.; Suzuki, T.; Yamano, S.; Gonzalez, F.J. Hepatic steatosis in leptin-deficient mice is promoted by the PPARgamma target gene Fsp27. Cell Metab. 2008, 7, 302–311. [Google Scholar] [CrossRef] [PubMed]
  119. Ilchovska, D.D.; Barrow, D.M. An Overview of the NF-kB mechanism of pathophysiology in rheumatoid arthritis, investigation of the NF-kB ligand RANKL and related nutritional interventions. Autoimmun. Rev. 2021, 20, 102741. [Google Scholar] [CrossRef] [PubMed]
  120. Creed, T.J.; Probert, C.S. Review article: Steroid resistance in inflammatory bowel disease—Mechanisms and therapeutic strategies. Aliment. Pharm. 2007, 25, 111–122. [Google Scholar] [CrossRef]
  121. Sohrabi, M.; Gholami, A.; Amir Kalali, B.; Khoonsari, M.; Sahraei, R.; Nasiri Toosi, M.; Zamani, F.; Keyvani, H. Are Serum Levels of Nuclear Factor Kappa B and Forkhead Box Protein P3 in Patients with Non-Alcoholic Fatty Liver Disease Related to Severity of Fibrosis? Middle East J. Dig. Dis. 2021, 13, 356–362. [Google Scholar] [CrossRef]
  122. Xu, X.; Wang, W.; Lin, L.; Chen, P. Liraglutide in combination with human umbilical cord mesenchymal stem cell could improve liver lesions by modulating TLR4/NF-kB inflammatory pathway and oxidative stress in T2DM/NAFLD rats. Tissue Cell 2020, 66, 101382. [Google Scholar] [CrossRef]
  123. Mehta, M.; Duseja, A.; Gupta, S. Phyllanthin Regulates Expression of NFKB/PI3K/AKT Pathway and Improves Liver Histology in Animal Model of NAFLD: A Preliminary Study. J. Clin. Exp. Hepatol. 2022, 12, S7. [Google Scholar] [CrossRef]
  124. Peluso, I.; Yarla, N.S.; Ambra, R.; Pastore, G.; Perry, G. MAPK signalling pathway in cancers: Olive products as cancer preventive and therapeutic agents. Semin. Cancer Biol. 2019, 56, 185–195. [Google Scholar] [CrossRef]
  125. Wu, L.; Liu, Y.; Zhao, Y.; Li, M.; Guo, L. Targeting DUSP7 signaling alleviates hepatic steatosis, inflammation and oxidative stress in high fat diet (HFD)-fed mice via suppression of TAK1. Free Radic. Biol. Med. 2020, 153, 140–158. [Google Scholar] [CrossRef] [PubMed]
  126. Nitulescu, G.M.; Van De Venter, M.; Nitulescu, G.; Ungurianu, A.; Juzenas, P.; Peng, Q.; Olaru, O.T.; Gradinaru, D.; Tsatsakis, A.; Tsoukalas, D.; et al. The Akt pathway in oncology therapy and beyond (Review). Int. J. Oncol. 2018, 53, 2319–2331. [Google Scholar] [CrossRef] [PubMed]
  127. Huang, X.; Liu, G.; Guo, J.; Su, Z. The PI3K/AKT pathway in obesity and type 2 diabetes. Int. J. Biol. Sci. 2018, 14, 1483–1496. [Google Scholar] [CrossRef] [PubMed]
  128. Yuan, Y.; Ma, M.; Zhang, S. Recent advances in delivery systems of fucoxanthin. Food Chem. 2023, 404, 134685. [Google Scholar] [CrossRef] [PubMed]
  129. Saravana, P.S.; Shanmugapriya, K.; Gereniu, C.R.N.; Chae, S.-J.; Kang, H.W.; Woo, H.-C.; Chun, B.-S. Ultrasound-mediated fucoxanthin rich oil nanoemulsions stabilized by κ-carrageenan: Process optimization, bio-accessibility and cytotoxicity. Ultrason. Sonochemistry 2019, 55, 105–116. [Google Scholar] [CrossRef]
  130. Sun, X.; Xu, Y.; Zhao, L.; Yan, H.; Wang, S.; Wang, D. The stability and bioaccessibility of fucoxanthin in spray-dried microcapsules based on various biopolymers. RSC Adv. 2018, 8, 35139–35149. [Google Scholar] [CrossRef] [PubMed]
  131. Yuan, Y.; Ma, M.; Wang, D.; Xu, Y. A review of factors affecting the stability of zein-based nanoparticles loaded with bioactive compounds: From construction to application. Crit. Rev. Food Sci. Nutr. 2022, 1–17, Epub ahead of print. [Google Scholar] [CrossRef]
  132. Li, Y.; Dou, X.; Pang, J.; Liang, M.; Feng, C.; Kong, M.; Liu, Y.; Cheng, X.; Wang, Y.; Chen, X. Improvement of fucoxanthin oral efficacy via vehicles based on gum Arabic, gelatin and alginate hydrogel: Delivery system for oral efficacy enhancement of functional food ingredients. J. Funct. Foods 2019, 63, 103573. [Google Scholar] [CrossRef]
  133. Wang, C.; Ren, J.; Song, H.; Chen, X.; Qi, H. Characterization of whey protein-based nanocomplex to load fucoxanthin and the mechanism of action on glial cells PC12. Lwt 2021, 151, 112208. [Google Scholar] [CrossRef]
  134. Wang, L.; Wei, Z.; Xue, C.; Tang, Q.; Zhang, T.; Chang, Y.; Wang, Y. Fucoxanthin-loaded nanoparticles composed of gliadin and chondroitin sulfate: Synthesis, characterization and stability. Food Chem. 2022, 379, 132163. [Google Scholar] [CrossRef] [PubMed]
  135. Liu, Y.; Qiao, Z.; Liu, W.; Hou, Z.; Zhang, D.; Huang, L.; Zhang, Y. Oleic acid as a protein ligand improving intestinal absorption and ocular benefit of fucoxanthin in water through protein-based encapsulation. Food Funct. 2019, 10, 4381–4395. [Google Scholar] [CrossRef]
  136. Malgarim Cordenonsi, L.; Faccendini, A.; Catanzaro, M.; Bonferoni, M.C.; Rossi, S.; Malavasi, L.; Platcheck Raffin, R.; Scherman Schapoval, E.E.; Lanni, C.; Sandri, G.; et al. The role of chitosan as coating material for nanostructured lipid carriers for skin delivery of fucoxanthin. Int. J. Pharm. 2019, 567, 118487. [Google Scholar] [CrossRef] [PubMed]
  137. Zhu, J.; Li, H.; Xu, Y.; Wang, D. Construction of fucoxanthin vector based on binding of whey protein isolate and its subsequent complex coacervation with lysozyme. J. Agric. Food Chem. 2019, 67, 2980–2990. [Google Scholar] [CrossRef]
  138. Albracht-Schulte, K.; Kalupahana, N.S.; Ramalingam, L.; Wang, S.; Rahman, S.M.; Robert-McComb, J.; Moustaid-Moussa, N. Omega-3 fatty acids in obesity and metabolic syndrome: A mechanistic update. J. Nutr. Biochem. 2018, 58, 1–16. [Google Scholar] [CrossRef]
  139. Kersten, S.; Mandard, S.; Tan, N.S.; Escher, P.; Metzger, D.; Chambon, P.; Gonzalez, F.J.; Desvergne, B.; Wahli, W. Characterization of the fasting-induced adipose factor FIAF, a novel peroxisome proliferator-activated receptor target gene. J. Biol. Chem. 2000, 275, 28488–28493. [Google Scholar] [CrossRef] [PubMed]
  140. Chen, L.; Duan, Y.; Wei, H.; Ning, H.; Bi, C.; Zhao, Y.; Qin, Y.; Li, Y. Acetyl-CoA carboxylase (ACC) as a therapeutic target for metabolic syndrome and recent developments in ACC1/2 inhibitors. Expert Opin. Investig. Drugs 2019, 28, 917–930. [Google Scholar] [CrossRef]
  141. Qian, G.; Morral, N. Role of non-coding RNAs on liver metabolism and NAFLD pathogenesis. Hum. Mol. Genet. 2022, 31, R4–R21. [Google Scholar] [CrossRef]
  142. Shen, X.; Guo, H.; Xu, J.; Wang, J. Inhibition of lncRNA HULC improves hepatic fibrosis and hepatocyte apoptosis by inhibiting the MAPK signaling pathway in rats with nonalcoholic fatty liver disease. J. Cell Physiol. 2019, 234, 18169–18179. [Google Scholar] [CrossRef]
  143. Zhang, Z.; Liu, X.; Xu, H.; Feng, X.; Lin, Y.; Huang, Y.; Peng, Y.; Gu, M. Obesity-induced upregulation of miR-361-5p promotes hepatosteatosis through targeting Sirt1. Metabolism 2018, 88, 31–39. [Google Scholar] [CrossRef] [PubMed]
  144. Meng, X.; Guo, J.; Fang, W.; Dou, L.; Li, M.; Huang, X.; Zhou, S.; Man, Y.; Tang, W.; Yu, L.; et al. Liver MicroRNA-291b-3p Promotes Hepatic Lipogenesis through Negative Regulation of Adenosine 5’-Monophosphate (AMP)-activated Protein Kinase alpha1. J. Biol. Chem. 2016, 291, 10625–10634. [Google Scholar] [CrossRef]
  145. Liu, W.; Cao, H.; Ye, C.; Chang, C.; Lu, M.; Jing, Y.; Zhang, D.; Yao, X.; Duan, Z.; Xia, H.; et al. Hepatic miR-378 targets p110alpha and controls glucose and lipid homeostasis by modulating hepatic insulin signalling. Nat. Commun. 2014, 5, 5684. [Google Scholar] [CrossRef]
  146. Sun, X.; Zhao, H.; Liu, Z.; Sun, X.; Zhang, D.; Wang, S.; Xu, Y.; Zhang, G.; Wang, D. Modulation of Gut Microbiota by Fucoxanthin During Alleviation of Obesity in High-Fat Diet-Fed Mice. J. Agric. Food Chem. 2020, 68, 5118–5128. [Google Scholar] [CrossRef] [PubMed]
  147. Guo, B.; Yang, B.; Pang, X.; Chen, T.; Chen, F.; Cheng, K.-W. Fucoxanthin modulates cecal and fecal microbiota differently based on diet. Food Funct. 2019, 10, 5644–5655. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Fucoxanthin and its metabolites. (A) Fucoxanthin, (B) fucoxanthinol, and (C) amarouciaxanthin A.
Figure 1. Fucoxanthin and its metabolites. (A) Fucoxanthin, (B) fucoxanthinol, and (C) amarouciaxanthin A.
Ijms 24 08203 g001
Figure 2. The role of fucoxanthin against NAFLD. Fucoxanthin exhibits anti-inflammatory, anti-obesity, anti-oxidant, and anti-fibrogenic activities and alters lipid metabolism in NAFLD.
Figure 2. The role of fucoxanthin against NAFLD. Fucoxanthin exhibits anti-inflammatory, anti-obesity, anti-oxidant, and anti-fibrogenic activities and alters lipid metabolism in NAFLD.
Ijms 24 08203 g002
Figure 3. Fucoxanthin-induced thermogenic activity. Fucoxanthin induces the activation of UCP-1, leading to browning of white adipose tissues. Uncoupled proton leak will occur instead of coupling with oxidative phosphorylation along with heat release (adapted from Huang, (2022)) [50].
Figure 3. Fucoxanthin-induced thermogenic activity. Fucoxanthin induces the activation of UCP-1, leading to browning of white adipose tissues. Uncoupled proton leak will occur instead of coupling with oxidative phosphorylation along with heat release (adapted from Huang, (2022)) [50].
Ijms 24 08203 g003
Figure 4. Fucoxanthin alters lipid metabolism. (A) Lipogenesis. Fucoxanthin downregulates acetyl CoA carboxylase and fatty acid synthase in lipogenesis, decreasing the lipid content in the liver. (B) Lipolysis. Fucoxanthin promotes β-oxidation, hence resulting in upregulated lipolysis in the liver. The downward arrow indicates downregulation and the upward arrow indicates upregulation.
Figure 4. Fucoxanthin alters lipid metabolism. (A) Lipogenesis. Fucoxanthin downregulates acetyl CoA carboxylase and fatty acid synthase in lipogenesis, decreasing the lipid content in the liver. (B) Lipolysis. Fucoxanthin promotes β-oxidation, hence resulting in upregulated lipolysis in the liver. The downward arrow indicates downregulation and the upward arrow indicates upregulation.
Ijms 24 08203 g004
Figure 5. Signaling pathways altered by fucoxanthin. Fucoxanthin alters β-3-adrenergic receptor (β3Ad), adenosine monophosphate-activated protein kinase (AMPK), peroxisome proliferator-activated receptor (PPAR), sterol regulatory element binding protein (SREBP), nuclear factor-κB (NF-κB), mitogen-activated protein kinase (MAPK), and AKT pathways.
Figure 5. Signaling pathways altered by fucoxanthin. Fucoxanthin alters β-3-adrenergic receptor (β3Ad), adenosine monophosphate-activated protein kinase (AMPK), peroxisome proliferator-activated receptor (PPAR), sterol regulatory element binding protein (SREBP), nuclear factor-κB (NF-κB), mitogen-activated protein kinase (MAPK), and AKT pathways.
Ijms 24 08203 g005
Table 2. Transcriptional factors altered by fucoxanthin.
Table 2. Transcriptional factors altered by fucoxanthin.
Transcriptional FactorsExperimental ModelDoseReference(s)
Thermogenesis Related
↑ PGC-1KK-Ay miceUndaria pinnatifida-extracted fucoxanthin; 0.5% and 2% of control diet;[51]
0.2% fucoxanthin in control diet (AIN-93G)[105]
High-fat-diet-induced obese mice0.69% Undaria pinnatifida-extracted fucoxanthin (2.9%)[76]
↑ β3-adrenergic receptor (β3Ad)KK-Ay miceUndaria pinnatifida-extracted fucoxanthin; 0.5% and 2% of control diet;[51]
High-fat-diet-induced obese mice1.06% or 2.22% in control diet (AIN-93G)[57]
Lipid Metabolism Related
↑ AMPKHigh-fat-diet-induced obese micePetalonia binghamiae-extracted fucoxanthin; 150 mg/kg/day[7]
↑ PPARαUndaria pinnatifida-extracted fucoxanthin (2.9%); 0.69% w/w[76]
↑ PPAR β0.05% and 0.2% fucoxanthin, w/w[9]
↓ PPAR γHigh-fat-diet-induced obese mice0.05% and 0.2% fucoxanthin, w/w[9]
Petalonia binghamiae-extracted fucoxanthin; 150 mg/kg/day[7]
Undaria pinnatifida-extracted fucoxanthin (2.9%); 0.69% w/w[76]
3T3-L1Petalonia binghamiae-extracted fucoxanthin; 10 μM[106]
fucoxanthin, fucoxanthinol, and amarouciaxanthin extracted from U. pinnatifida; 10 μM[107]
↓ SREBP1cHigh-fat-diet-induced obese micePetalonia binghamiae-extracted fucoxanthin; 150 mg/kg/day[7]
Inflammation Related
↓ NF-κBRAW 264.7 macrophagesI. okamurae-extracted fucoxanthin; 12.5, 25, or 50 μM[86]
Carr-induced paw edema in ICR miceUndaria pinnatifida-extracted fucoxanthin; 4 and 8 mg/kg[89]
↓ MAPKMacrophage RAW 264.7 cellsI. okamurae-extracted fucoxanthin; 12.5, 25, or 50 μM[86]
Carr-induced paw edema in ICR miceUndaria pinnatifida-extracted fucoxanthin; 4 and 8 mg/kg[89]
↓ Akt
Anti-Oxidant Related
↑ Nrf2Alcoholic liver injury mouse model10, 20, 40 mg/kg b.w.[108]
H9c2 cells1 μM[109]
↑ AMPKHepC2 cellsL. Japonica-extracted fucoxanthin; 30 μg/mL[110]
Anti-Fibrogenic
↓ SMAD2/3Human pulmonary fibroblasts (HPFs)5, 10, 20 μM[111]
↓ PI3K/Akt
↓ MAPK
↑—upregulated; ↓—downregulated.
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

Winarto, J.; Song, D.-G.; Pan, C.-H. The Role of Fucoxanthin in Non-Alcoholic Fatty Liver Disease. Int. J. Mol. Sci. 2023, 24, 8203. https://doi.org/10.3390/ijms24098203

AMA Style

Winarto J, Song D-G, Pan C-H. The Role of Fucoxanthin in Non-Alcoholic Fatty Liver Disease. International Journal of Molecular Sciences. 2023; 24(9):8203. https://doi.org/10.3390/ijms24098203

Chicago/Turabian Style

Winarto, Jessica, Dae-Geun Song, and Cheol-Ho Pan. 2023. "The Role of Fucoxanthin in Non-Alcoholic Fatty Liver Disease" International Journal of Molecular Sciences 24, no. 9: 8203. https://doi.org/10.3390/ijms24098203

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