Next Article in Journal
Deconstructing the Supermarket: Systematic Ingredient Disaggregation and the Association between Ingredient Usage and Product Health Indicators for 24,229 Australian Foods and Beverages
Previous Article in Journal
Association between Food Insecurity and Sarcopenia among Adults Aged ≥65 Years in Low- and Middle-Income Countries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Products in the Prevention of Metabolic Diseases: Lessons Learned from the 20th KAST Frontier Scientists Workshop

by
Seung J. Baek
1,
Bruce D. Hammock
2,
In-Koo Hwang
1,
Qing X. Li
3,
Naima Moustaid-Moussa
4,
Yeonhwa Park
5,
Stephen Safe
6,
Nanjoo Suh
7,
Sun-Shin Yi
8,
Darryl C. Zeldin
9,
Qixin Zhong
10,
Jennifer Alyce Bradbury
9,
Matthew L. Edin
9,
Joan P. Graves
9,
Hyo-Young Jung
1,
Young-Hyun Jung
1,
Mi-Bo Kim
11,
Woosuk Kim
1,
Jaehak Lee
1,
Hong Li
9,
Jong-Seok Moon
8,
Ik-Dong Yoo
8,
Yiren Yue
5,
Ji-Young Lee
11,* and
Ho-Jae Han
1,*
add Show full author list remove Hide full author list
1
College of Veterinary Medicine, Seoul National University, Seoul 08826, Korea
2
Department of Entomology, University of California, Davis, CA 95616, USA
3
Department of Molecular Biosciences and Bioengineering, University of Hawaii at Manoa, Honolulu, HI 96822, USA
4
Department of Nutritional Sciences & Obesity Research Institute, Texas Tech University, Lubbock, TX 79409, USA
5
Department of Food Science, University of Massachusetts, Amherst, MA 01003, USA
6
Department of Biochemistry & Biophysics, Texas A & M University, College Station, TX 77843, USA
7
Department of Chemical Biology, Ernest Mario School of Pharmacy, Rutgers University, Piscataway, NJ 08854, USA
8
Department of Medical Sciences, Soonchunhyang University, Asan 31538, Korea
9
National Institutes of Environmental Health, National Institutes of Health, Research Triangle Park, NC 27709, USA
10
Department of Food Sciences, University of Tennessee, Knoxville, TN 37996, USA
11
Department of Nutritional Sciences, University of Connecticut, Storrs, CT 06269, USA
*
Authors to whom correspondence should be addressed.
Nutrients 2021, 13(6), 1881; https://doi.org/10.3390/nu13061881
Submission received: 17 March 2021 / Revised: 20 May 2021 / Accepted: 20 May 2021 / Published: 31 May 2021
(This article belongs to the Section Phytochemicals and Human Health)

Abstract

:
The incidence of metabolic and chronic diseases including cancer, obesity, inflammation-related diseases sharply increased in the 21st century. Major underlying causes for these diseases are inflammation and oxidative stress. Accordingly, natural products and their bioactive components are obvious therapeutic agents for these diseases, given their antioxidant and anti-inflammatory properties. Research in this area has been significantly expanded to include chemical identification of these compounds using advanced analytical techniques, determining their mechanism of action, food fortification and supplement development, and enhancing their bioavailability and bioactivity using nanotechnology. These timely topics were discussed at the 20th Frontier Scientists Workshop sponsored by the Korean Academy of Science and Technology, held at the University of Hawaii at Manoa on 23 November 2019. Scientists from South Korea and the U.S. shared their recent research under the overarching theme of Bioactive Compounds, Nanoparticles, and Disease Prevention. This review summarizes presentations at the workshop to provide current knowledge of the role of natural products in the prevention and treatment of metabolic diseases.

1. Introduction

Bioactive food components, particularly phytochemicals, provide various health benefits in experimental animals and humans. As our knowledge grows regarding the pathogenesis of obesity-associated diseases, neurological diseases, and aging, identifying bioactive compounds that exert antioxidant and anti-inflammatory properties from natural products has been a crucial avenue to advance biomedical, nutrition, and food research.
Oxidative stress from the excessive production of reactive oxygen species (ROS) or free radicals in cells leads to an imbalance in cellular homeostasis and damage in DNA, proteins and lipids, further contributing to chronic inflammation [1,2,3]. Growing evidence indicates that oxidative stress and inflammation are closely linked as key drivers in the pathogenesis of many chronic diseases, including cancers, obesity, diabetes, cardiovascular and neurological diseases [3,4,5,6]. Notably, many experimental and epidemiological studies suggest that consuming fruits, vegetables, nuts, spices, whole grains, and plant-based oils may reduce inflammation and oxidative stress [7]. Dietary components, such as flavonoids, catechins, curcumin, resveratrol, sulforaphane, silymarin, and xantho-humol, are known to exert anti-inflammatory and antioxidant properties [7]. Short-term or acute consumption of these bioactive components is not effective in reducing persisting oxidative stress or chronic inflammatory conditions. Healthy lifestyle changes, such as exercise and healthy dietary habits, are more effective approaches to prevent chronic inflammatory conditions, cancers, and neurological diseases [8,9].
This review provides a current understanding of key bioactive food compounds and their mode of action, and innovative tools developed to identify putative health-promoting compounds. The ultimate goal is to incorporate these bioactives into the food system for optimal health. These timely topics were discussed at the 20th Frontier Scientists Workshop sponsored by the Korean Academy of Science and Technology.

2. Inflammation, Oxidative Stress, and Natural Products

2.1. Antioxidant Effect of Triterpenoids and Tocopherols

Many dietary factors and phytochemicals have been shown to exhibit inhibitory effects on oxidative stress and inflammatory responses [10]. Key examples of bioactive compounds targeting oxidative stress and inflammation are triterpenoids [11] and tocopherols [12]. Several hundred oleanane triterpenoids have been synthesized based on anti-inflammatory assays [13,14,15]. Many of these triterpenoids suppress inflammation and oxidative stress and exert cytoprotective effects via activation of nuclear factor erythroid 2-related factor 2 (NRF2), a master transcription factor in the antioxidant defense systems [16]. We have shown that tocopherols and tocotrienols are potential natural antioxidant and anti-inflammatory agents that may be used as dietary cancer preventive agents. Tocopherols (Ts) and tocotrienols (T3s) are members of the vitamin E family, each form consisting of a chromanol ring and a side chain containing 16-carbons [17,18]. Depending on the number and position of the methyl group on the ring, they exist as α, β, δ, γTs or T3s [19,20]. Ts or T3s are phenolic antioxidants present in many vegetable oils, such as soybean, corn, canola and cottonseeds [17,21]. Vitamin Es are essential lipid-soluble antioxidants, scavenging reactive oxygen/nitrogen radicals in lipid milieu, and therefore they protect membrane integrity. They also inhibit inflammatory responses and other disorders [17,22,23]. T3s, having similar structures except for three double bonds on the side chains, also possess vitamin E activity [12].
Epidemiological studies have shown an inverse association between dietary intake of tocopherols or blood levels of tocopherols and cancer risk [18,24]. Most previous cancer prevention studies on vitamin E have focused on αT, the commonly recognized vitamin E form [18,23,25,26,27]. The most recent large clinical trial, the Selenium and Vitamin E Cancer Prevention Trial (SELECT) with αT, did not demonstrate a prostate cancer preventive effect [28,29]. The reasons for this result are not known, reflecting a lack of understanding of the biological activities of individual tocopherols. However, recent laboratory studies, including our studies with animal models for breast and other cancers [30,31,32,33,34,35,36,37], have demonstrated that γT and δT are more active than αT in inhibiting carcinogenesis. αT has methyl groups in both positions, and γT has only one methyl group adjacent to the phenolic group. γT is a more effective scavenger of reactive nitrogen species and a more effective inhibitor of oxidation of phospholipids than αT [38,39,40,41]. Due to the the fact γT has stronger anti-nitrosative and anti-inflammatory activities, its preventive actions against cancer, cardiovascular and neurodegenerative diseases have been suggested [38,40,42,43,44,45].

2.2. Role of Tart Cherry (TC) in the Prevention of Obesity-Related Inflammation and Life Span Extension

Obesity is a complex disease and a major health problem worldwide. A primary underlying etiology of obesity is chronic low-grade inflammation, triggered in part by adipose tissue expansion. This expansion leads to endocrine dysfunctions of adipose tissue, including local inflammation, resulting in low-grade systemic inflammation [46]. Obesity and aging share several metabolic, cell, and molecular dysfunctions. Commonly altered processes include inflammation, oxidative stress, and cellular damage, which accelerate aging [47]. Several cell culture, preclinical and clinical studies demonstrated that foods containing bioactive compounds exert protective effects in both obesity and aging-related inflammation and oxidative stress [48]. A few studies have focused on the benefits of consuming anthocyanin-rich food, such as reducing chronic metabolic diseases and improving age-related disorders [49,50]. Increased adipose tissue mass during obesity is linked closely with adipose tissue inflammation, a major contributor to other metabolic disorders, including cardiovascular disease and type 2 diabetes [46], as documented in several diet-induced and genetic models of obesity [46,51,52].
Tart cherry (TC) (Prunus cerasus) contains higher levels of anthocyanin flavonoids than other cherries [53]. TC supplementation reduced adipose tissue inflammation, as demonstrated by the downregulation of several pro-inflammatory cytokines, such as IL-6, TNFα, IL-1β, MCP-1, iNOS, and CD-11b, and increased gene expression of anti-inflammatory M2 macrophage markers in adipose tissue of genetically obese Zucker rats [52]. In cultured adipocytes, TC extracts reduced LPS-induced expression of pro-inflammatory markers. The anti-inflammatory effects of TC are, in part, mediated by the inhibition nuclear factor κB (NFκB) p65 phosphorylation, both in vivo and in vitro [52]. TC also decreased expression of lipogenic markers and increased expression of lipid oxidizing and antioxidant genes both in vivo and in vitro, resulting in reduced lipid accumulation in adipocytes. These findings are consistent with other reports that regular intake of TC and other anthocyanin-rich fruits reduced systemic and local inflammation, including adipose tissue inflammation in obesity [54,55,56]. Several clinical studies also showed that increased intake of fruits and vegetables reduced markers of inflammation and oxidative stress in adults [57]. This is presumably linked to a high content of antioxidant and anti-inflammatory compounds in fruits and vegetables. Thus, TC rich in anthocyanins may be a potent anti-obesity and dietary anti-inflammatory component. However, further research on how TC affects adipose tissue and energy metabolism is needed. Especially, clinical studies are lacking to determine an optimal beneficial dosage for human consumption to prevent or reduce obesity and inflammation.
Obesity affects normal cellular and molecular processes, which can reduce lifespan expectancy [47]. Dietary antioxidants have potential roles in lowering oxidative stress associated with aging and chronic conditions [58]. We have recently reported that TC extract rich in anthocyanins with antioxidant activity increased lifespan in C. elegans mainly via insulin/insulin-like growth factor-1 signaling (IIS) by regulating DAF-2 and DAF-16 expression, major components of the IIS pathway [59]. Changes during aging are often linked to mitochondrial dysfunction. For the first time, we demonstrated that worms fed TC extracts exhibited increased mitochondrial spare respiration, and expression of uncoupling protein 4 and antioxidant markers such as superoxide dismutase (SOD)-3 [59]. Additional research is warranted both in mouse models and clinical studies to better understand the detailed molecular mechanisms of TC in aging and determine the most effective dose for human consumption.

2.3. Food-Derived Antioxidants and Lifespan

Oxidative stress contributes to the development of a range of adverse health conditions, including aging and age-related diseases [60]. Caenorhabditis elegans is a multi-organ, microscopic, and transparent roundworm used in many scientific research fields [61]. Particularly, aging research has taken advantage of its short lifespan and simple physiology in the recent decades [62]. Along with genetic manipulation, more environmental manipulative approaches have recently been used in aging studies with C. elegans, such as screening for antioxidant and anti-aging food bioactives, due to the well-conserved aging and stress-related pathways in this organism [61,62].
Using C. elegans as a model, we determined the effect of piceatannol, chicoric acid and p-coumaric acid, phenolic compounds widely found in plant food [63,64,65], on lifespan extension. All three compounds displayed strong antioxidant properties with a 15%–38% reduction in internal ROS levels in C. elegans [66,67,68]. In addition, they improved the worms’ survival under the paraquat-induced oxidative stress condition [66,67,68]. As a hydroxylated resveratrol derivative, piceatannol regulated oxidative stress responses via sir-2.1 (encodes a homolog of NAD-dependent deacetylase sirtuin-1, SIRT1) and daf-16 (encodes a homolog of Forkhead box O transcription factor, FoxO) [66], which have previously been suggested as resveratrol’s targets [69,70]. Chicoric acid and p-coumaric acid, hydroxyl derivatives of cinnamic acid, shared the same target, skn-1 (encoding a homolog of NRF2). Chicoric acid also activated aak-2 (encoding a homolog of AMP-activated protein kinase α) [67], which further contributed to its antioxidative activities.
Although piceatannol, chicoric acid, and p-coumaric acid exhibited potent antioxidant properties in C. elegans, their lifespan extension capacities were not the same. Piceatannol [66] and chicoric acid [67] significantly increased the mean lifespan by ~18% and ~24%, respectively, compared to the control, while p-coumaric acid extended the lifespan only under oxidative stress conditions [68]. These observations suggest that the antioxidative property of the compounds is not a sole contributor to an extended lifespan. Similar observations of enhanced oxidative stress responses without lifespan extension were also reported for other food bioactive/ingredients [71,72,73].
Despite its several advantages, C. elegans also has limitations, such as a lack of particular organs and circulatory system [74]. Therefore, further investigations are needed to evaluate the effects of food bioactives on aging and age-related diseases in C. elegans using vertebrate animals and, eventually, humans [75].

2.4. Role of Phytochemicals in the Regulation of Mitochondrial Functions under Oxidative Stress

The therapeutic effects of stem cells are well-known in a clinical setting. Mesenchymal stem cells (MSCs) and their secretory factors have been extensively used to develop therapeutic drugs targeting tissue regeneration, anti-inflammation, and immune modulation [76]. However, there are some limitations. MSCs cannot be produced indefinitely due to their limited proliferation and replication capacity [77,78]. Furthermore, transplanted stem cells exposed to a low-oxygen environment in target organs or bloodstream do not function effectively due to reduced survival rate, differentiation potential, and proliferation [79,80]. MSCs exposed to oxidative stress trigger a cell protection mechanism known as hypoxic adaptation. Recent studies have suggested that hypoxic adaptation is closely related to mitochondria function vital to maintaining stem cell self-renewal ability [81]. Thus, using antioxidant bioactive molecules is a promising approach to help stem cells adapt to oxidative stress, ultimately improving their therapeutic efficacy.
Phytochemicals, such as ascorbic acid, carotenoids, phenolic compounds, flavonoids, and terpenoids, have potent antioxidant and anti-inflammatory effects [82]. Studies of the regulatory effects of phytochemicals on mitochondria function have been limited to their ROS-scavenging properties. However, it has emerged that phytochemicals may play crucial roles in the regulation of MSC proliferation and differentiation by maintaining mitochondrial functions in oxidative stress conditions [83]. Preconditioning of MSCs under hypoxic conditions enhances their therapeutic effects via metabolic alterations in mitochondrial functions [84]. Primary targeting functions related to mitochondrial physiology during metabolic alteration include excessive mitochondrial respiration, accumulation of mitochondrial ROS, altered mitochondrial dynamics, and mitophagy inhibition [85,86]. We found that BCL2/adenovirus E1B 19 KDa protein-interacting protein 3 (BNIP3) is a major mitophagy regulatory protein induced by hypoxia in MSCs, contributing to sustaining the therapeutic function of MSCs by maintaining mitochondrial ROS and membrane potential homeostasis [87]. Interestingly, hypoxia-induced factor 1 α (HIF1α)-dependent downregulation of BNIP3 under high-glucose was rescued by tetra-methylpyrazine, an alkyl-pyrazine found in fermented cocoa beans [88].
Recent studies have also suggested that dietary phytochemicals, such as resveratrol, curcumin, and sulforaphane, have protective effects against mitochondrial dysfunction [89,90]. The SIRT family is the primary target protein of resveratrol, a phytoalexin present in fruits, in response to injury or infection. Resveratrol-activated SIRT1 induced SOD expression and rescued apoptosis, lowering ROS levels in ischemic diseases [91]. Our previous studies consistently showed that activation of SIRT3 reduces mitochondrial ROS and maintains mitochondrial functions under oxidative stress, ameliorating apoptosis and replicative senescence of MSCs [92,93]. Fucoidan, a sulfated polysaccharide found in the fibrillar cell walls and intercellular spaces of brown seaweed, also enhanced the therapeutic effects of MSCs in ischemic disease [94,95]. Importantly, fucoidan rescued ischemia-induced MSC apoptosis by the induction of SOD-2 to scavenge mitochondrial ROS, suggesting that fucoidan may be used for clinical application of stem cell therapy [96]. We also showed that O-cyclic phytosphingosine-1-phosphate (cP1P), a phytochemical-derived novel molecule, induced glycolytic metabolic alterations and suppressed mitochondrial ROS accumulation via HIF1α nuclear translocation in an oxidative stress condition [84]. Therefore, cP1P has great potential as a scalable bioactive molecule that can enhance the therapeutic efficacy of MSCs for the treatment of various diseases.

2.5. Role of Cyclooxygenases in T Cell Differentiation and Function

Cyclooxygenase-1 and cyclooxygenase-2 (COX-1 and COX-2, encoded by the PTGS1 and PTGS2 genes, respectively) metabolize arachidonic acid to prostaglandin (PG) H2, which is the substrate for synthases that generate specific PGs with varied and potent biological effects [97]. COX-1 is constitutively expressed in most cell types and considered the “housekeeping” isoform that produces PGs involved in maintaining normal cell/organ function. In contrast, COX-2 is induced by pro-inflammatory cytokines, mitogens and other environmental stimuli, e.g., LPS, allergens, and produces PGs involved in regulating inflammatory processes [97]. The COX enzymes are major targets of the widely prescribed non-steroidal anti-inflammatory drugs (NSAIDs), e.g., ibuprofen, aspirin, and selective COX-2 inhibitors, e.g., celecoxib, which reduce prostanoid biosynthesis [98]. PG biosynthesis is altered in individuals with SNPs in the PTGS1/2 isoforms [99,100,101]. While the contributions of COX-derived eicosanoids to the pathogenesis of cancer and cardiovascular disease are well documented, their role in the immune system in general and T cell function, in particular, has been under-investigated.
COX-derived prostaglandins can regulate the differentiation of CD4+ T helper (Th) cells to Th1, Th2, Th9 and Th17 subsets, which are key components of the adaptive immune response. Dendritic cells (DCs) present allergens to naïve CD4+ T cells and produce immunomodulatory cytokines that induce differentiation to specific Th cell subsets. Thromboxane A2 (TXA2) reduces the strength of DC/T cell interactions and suppresses T cell differentiation [102]. PGE2 and PGI2 reduce DC-derived IL-12 formation and subsequent Th1 differentiation [103,104]. We previously demonstrated that COX-1−/− mice have an accentuated Th2 airway response after allergen challenge [105]. More recently, we found impaired Th17 cell differentiation of COX-2−/− naïve CD4+ T cells with decreased Stat3 phosphorylation and RORγ expression [106]. Synthetic PGF and PGI2 enhanced Th17 cell differentiation in vitro. In contrast, selective COX-2 inhibition or siRNA knockdown of the PGF (FP) or PGI2 (IP) receptors decreased Th17 cell differentiation in vitro. Importantly, the administration of synthetic PGs restored the accumulation of Th17 cells in the lungs of allergic COX-2−/− mice in vivo. Thus, COX-2 is a critical regulator of Th17 cell differentiation and function during allergic lung inflammation via autocrine signaling of PGI2 and PGF through their respective cell surface receptors [106]. Th9 cells promote inflammation by stimulating cytokine release and leukocyte proliferation and are involved in the pathogenesis of allergic asthma [107]. We found that COX-2 is a key negative regulator of Th9 cell differentiation and function in the allergic mouse lung [108]. In vitro studies revealed that both PGD2 and PGE2 suppress Th9 cell differentiation through activation of their cognate DP and EP receptors, enhanced PKA signaling, and suppression of IL-17RB and PU.1 expression. Pharmacologic treatment with PGD2 and PGE2 restored Th9 cell differentiation and function in COX-2−/− mice in vivo. Importantly, we found that human Th9 cell differentiation is also regulated by COX-2-derived eicosanoids.
Long before the discovery of COXs, members of ancient civilizations used natural COX inhibitors for medicinal purposes. The Babylonian, Chinese and Greek civilizations recognized the ability of poplar or willow bark extracts to reduce fever, inflammation, and pain [109]. These extracts were crude precursors to aspirin, which is widely used today. Aspirin acetylates and inactivates both COX-1 and COX-2 [110]. Other traditional herbal medicines are now known to include compounds that also act as COX inhibitors; these include alkaloids, terpenoids, stilbenes, flavonoids and saponins [111]. In addition, dietary regulation may impact inflammatory diseases. Omega-3 fatty acid rich diets contain eicosapentaenoic acid (EPA), which is metabolized to 3-series prostaglandins (e.g., PGE3, PGI3, TXA3) that are generally less inflammatory or have increased anti-inflammatory properties. For example, while PGE2 displays potent pro-inflammatory induction of both COX-2 and IL-6 in macrophages, PGE3-induced COX-2 and IL-6 induction is significantly lower [112].
In summary, COX enzymes and their eicosanoid products play a critical role in regulating immune cell function (Figure 1). Targeting cyclooxygenases and their eicosanoid products may represent a new approach to treating inflammatory diseases such as asthma [113].

3. Neuroinflammation and Natural Products

3.1. Protective Effects of Phytochemicals on Blood–Brain Barrier Integrity after Ischemic Damage to the Brain

Brain ischemia is one of the most critical neurological diseases. The majority of cases are associated with occlusion of blood vessels in the brain, leading to damage, disability, and reduction in patients’ quality of life [114,115]. The disruption of blood flow and reperfusion causes an imbalance in potassium and calcium ion levels in the extracellular fluid [116,117]. Calcium overload and subsequent high glutamate concentrations contribute to neuronal damage by inducing neuronal swelling [116]. Ischemia/reperfusion generates enormous amounts of ROS in the mitochondria, which increases the number of superoxide anions and hydroxyl radicals that induce DNA adducts and peroxidation of unsaturated fatty acids in the neuronal membranes [118]. Oxidative damage by ischemia/reperfusion causes the release of pro-inflammatory cytokines, and disruption in the integrity of the blood–brain barrier (BBB) facilitates the infiltration of leukocytes in the brain tissue after ischemia [119,120]. Restriction of brain inflammation by immune intervention reduces brain damage-induced ischemia [121]. Various studies have attempted to reduce the neuronal damage induced by ischemia because alteplase is the only acceptable drug by the FDA. However, the majority of the possible candidates showed some positive effects against ischemic damage.
The BBB consists of endothelial cells and basal lamina in blood vessels wrapped by foot processes of astrocytes and pericytes [122]. The endothelial cells are connected by tight junctions, which seal the BBB to control the transport of water and molecules across the barrier [122]. Ischemia causes impairment to the integrity of the BBB in a biphasic manner [123]. ROS produced by ischemia/reperfusion significantly increase the permeability transiently, which leads to partial recovery of the integrity of the BBB. Ischemia-induced inflammation causes a late breakdown of the BBB and results in infiltration of leukocytes in the brain tissue [124,125,126]. Transient forebrain ischemia activates matrix metalloproteinase (MMP)-2 and MMP-9, which cause chemokine-induced migration of leukocytes [127] in the ischemic penumbra and degrade the basal lamina in the blood vessels and endothelial tight junction proteins, leading to the breakdown of the BBB [128,129]. Activation of MMPs is tightly regulated by endogenous inhibitors of metalloproteinases (TIMPs), which bind to the active and alternative sites of MMPs [130,131]. TIMP-1 overexpression and synthetic MMP inhibitors reduce the disruption of the BBB and neuronal damage induced by ischemia in mice and rats [131,132]. Aquaporin-4 (AQP4) plays an important role in the transport of water in the brain, and its depletion in astrocytes significantly reduces water permeability [133,134], resulting in the amelioration of neuronal damage and brain edema after ischemia in AQP4-knockout mice [135]. Knockdown of FoxO3 significantly reduces hypoxia-induced hyperpermeability of the BBB and increases MMP-3 mRNA in bEnd.3 endothelial cells [136]. Vascular endothelial growth factor (VEGF) is one of the angiogenic factors in vascular endothelial cells, enhancing vascular permeability in the brain after ischemia [137,138].
Studies have investigated the therapeutic effects of phytochemicals against ischemic damage based on inhibition of disruption of the BBB (Table 1). Several phytochemicals show neuroprotective effects against ischemic damage by reducing MMP-9 and/or TIMP-1 levels in the BBB. In a human study, sodium tanshinone IIA sulfonate reduced the MMP-9, TIMP-1, and TIMP-2 levels in the serum of stroke patients who received recombinant tissue plasminogen activator (rt-PA) [139]. Treatment with ascorbic acid, chlorogenic acid, crocin, gastrodin, pinocembrin, and salvianolic acid A reduced BBB disruption by inhibiting MMP-2 activation in rats after ischemia [140,141,142,143,144,145,146]. Also, administration of astragaloside IV and ellagic acid reduced brain edema by AQP4 inhibition after ischemia in rats [147,148]. Dl-3-n-butylphthalide reduced the expression of caveolin-1, which contributes to tight junction disassembly and BBB leakage [149] after ischemia. Hesperidin demonstrated neuroprotective effects against hypoxic damage by decreasing FoxO3 and increasing MMP-3 mRNA levels in bEnd.3 endothelial cells [150]. Juglanin was shown to inhibit VEGF and its receptor (VEGFR2) signaling to improve the integrity of the BBB after ischemia [151]. Therefore, phytochemicals that reduce the permeability of the BBB can be used in various neurological disorders because many brain diseases are linked to the disruption of this barrier.

3.2. Fatty Acid Metabolites for the Treatment of Inflammation and Neuropathic Pain

The arachidonate cascade is where over 70% by mass of the world’s pharmaceuticals act by blocking two of its three branches, which are predominantly pro-inflammatory, leading to pain, hypertension, and other pathologies. Drugs acting on the arachidonate cascade include aspirin, indomethacin, ibuprofen, montelukast and others. We studied a third branch of the cascade termed the P450 branch, which produces natural fatty acid metabolites crucial for the resolution of inflammation and pain [164]. These natural metabolites called epoxy fatty acids (EpFA) are made by the P450 enzymes acting on polyunsaturated lipids from our diet. Our work has shown that these natural chemicals reduce the endoplasmic reticulum (ER) stress pathway, which is the underlying mechanism of diseases such as diabetes, lung fibrosis, heart failure, and even senescence and aging [165]. The ER stress pathway also can be activated by a variety of environmental contaminants.
After the EpFA are made by P450 or released from cellular membranes, they decrease the ER stress response and inflammation, thus reducing the severity of many diseases and symptoms such as pain. Although these natural product 1,2-disubstituted epoxides of fatty acids are chemically stable, EpFA are rapidly hydrolyzed to products that lack analgesic and pain-resolving properties and are somewhat pro-inflammatory. The hydrolysis is carried out by a second enzyme called the soluble epoxide hydrolase (sEH). The sEH enzyme is a homodimer of two α/β-fold monomers which converts the EpFA into their corresponding 1,2-diols. These diols have pro-inflammatory properties, are highly polar, and are rapidly conjugated and excreted. Thus, by inhibiting the sEH or by its genetic deletion, we can stabilize EpFA, increase their concentration, and stimulate resolution of pain and inflammation.
We have found that the sEH enzyme increases disease states and inflammation in the affected tissue, resulting in a decrease in EpFA. This, in turn, increases ER stress, triggering pain, inflammation, hypertension, and other disorders. We have prepared nanobodies to the sEH as a biomarker of inflammation and ER stress [166]. We have found that down-regulating or inhibiting the sEH can block the harmful actions of ER stress and also the toxicity of many environmental chemicals. Since EpFA are among the endothelium-derived hyperpolarizing factors, much of the work in the field has been on on vascular and renal inflammation, fibrosis, and heart diseases [167,168]. Recently we have looked increasingly at chronic neurological diseases [169]. With animal models or human stem cell models, we have shown that EpFA can prevent or even reverse such disorders as depression, autism, Parkinson’s disease, and Alzheimer’s disease (AD) [170,171,172,173,174]. Recent studies have indicated that sEH inhibitor (sEHI) can reduce colon tumorigenesis [175]. More broadly, the sEHI blocks the metastasis and tumor growth that often follows chemotherapy and resection [176]. These data suggest that NSAIDs could now be used to address this double-edged sword of cancer treatment enhancing tumor metastasis, followed by the more effective and safer sEHI. This fundamental knowledge allows us to understand disease states better as well as indicate what environmental compounds or combinations of compounds result in human risk.
From a translational perspective, we are introducing an injectable drug into the clinic to treat equine laminitis, which is a usually fatal disease characterized by intractable pain and inflammation, often in otherwise healthy horses. In parallel, we are providing the same compound in an oral formulation to treat inflammation in horses, dogs and cats with arthritis as an early indication [177,178]. We are taking forward a similar compound in a different patent class, the human market, with chronic or neuropathic pain as our major indication. We have investigational new drug approval from the US FDA through human efficacy testing and have completed human phase 1A safety trials with no adverse effects. Our specific clinical path is to address chronic pain with an oral compound, which is not an NSAID, has no addictive potential, and can spare or replace opioid analgesics [179]. Our recent research has demonstrated that the significant increase in plasma levels of regio-isomeric diols of linoleic acid formed by the sEH from the corresponding epoxides are by far the most significant predictors of severe outcomes of COVID-19 compared to over 120 other plasma biomarkers. These diols are termed leukotoxin diols and are both associated with and are potential causes of respiratory failure in humans, dogs, and mice. These pulmonary toxins may arise as an artifact of exceptionally high dietary linoleate in the western diet [180].

3.3. Role of Passiflora incarnata L. (PI) in Neuroinflammation

People, particularly in developed countries, have high rates of insomnia [181,182,183], disrupting the circadian rhythm. Exposure to light and darkness for an appropriate period significantly affects memory and learning by affecting hippocampal neurogenesis [184] and eating habits [185]. Lasting insomnia can cause severe neuroinflammation in the brain [186,187], causing clinical symptoms, such as abnormal eating behavior, depression, hormone secretion abnormalities, cerebrovascular diseases, cardiovascular diseases, and AD [188,189].
Natural products with minimal adverse side effects, including Passiflora incarnata L. (PI), one of the passion flowers, have recently received attention as an acute treatment for mild insomnia. PI has been widely known as a medicinal plant for anti-anxiety and sleep disorders in North America, Europe, Asia, Africa, and Australia for centuries. We found that the chronic administration of PI induces sleep quickly without any alterations in eating behaviors, body weight, and body composition while it significantly increases sleep duration compared to the vehicle-treated group. However, the identification of functional pharmacologic ingredients of PI has been limited. It is suggested that vitexin, one compound in PI, helps sleep and exerts anti-diabetic, anti-inflammatory and anti-AD actions [190,191,192,193,194,195,196,197]. We also confirmed that an ethanol extract of PI leaves and fruits has a high level of vitexin. A small dose of this PI extract led to high levels of hippocampal neurogenesis in rodent models without any behavioral and metabolic changes [190]. Although sleep was not induced, corticotropin-releasing factor (CRF) and glucocorticoid receptor (GR), pivotal regulatory factors for the hypothalamo–pituitary–adrenal (HPA) axis, were reduced, indicating its anti-stress effect. The vitexin and iso-vitexin present in PI have powerful anti-inflammatory and neuroprotective effects, which are expected to have potential pharmacologic effects for treating AD and other related diseases [192,198]. Our series of studies also strongly suggest that PI has potent anti-inflammatory and sleep-inducing effects.

4. Antioxidant and Anti-Inflammatory Compounds in Cancer Prevention

4.1. Roles of NAG-1 Activated by Phytochemicals in the Prevention of Inflammation-Mediated Tumorigenesis

Nonsteroidal anti-inflammatory drug (NSAID) activated gene-1 (NAG-1), also known as Growth Differentiation Factor 15 (GDF15), has been shown to have anti-inflammatory effects in NSAID-treated cells. In the last two decades, studies have demonstrated that NAG-1 plays a crucial role in various diseases, including inflammation, cancer, anorexia, cachexia, cardiovascular disease, neurodegenerative disease, diabetes, and obesity. Many anti-inflammatory phytochemicals are known to increase NAG-1 expression, preventing tumorigenesis.
Chronic inflammation can cause DNA damage in colon epithelial cells, which eventually turn into neoplastic lesions [199]. Several natural compounds have been reported to prevent colorectal cancer in vitro. 2′-Hydroxyflavanone, a compound rich in citrus fruits and vegetables, induced cell cycle arrest and apoptosis in HCT-116 cells, a human colorectal cancer cell line [200]. Apigenin, another naturally abundant molecule in fruits and vegetables, such as parsley, citrus fruits, and teas, exerted similar effects concomitantly with NAG-1 up-regulation promoted by PKCδ activation in HCT-116, LoVo, SW480, and HT-29 cells [201]. There was also an increase in NAG-1 expression with apoptosis when HCT-116 cells were treated with (−)-epigallocatechin gallate (EGCG) and (−)-epicatechin gallate (ECG), polyphenols in green tea [202], caffeic acid phenethyl ester (CAPE) [203], capsaicin [204], and isoliquiritin-genin, a chemical compound found in licorice [205]. Similarly, formononetin, commonly found in red clover and beans, exhibited apoptotic cell death, as evidenced by poly (ADP-ribose) polymerase cleavage, apoptotic body, and increased Annexin V-positive cells with NAG-1 upregulation in HCT-116 [205]. Pino-sylvin, a fungi-toxicant that protects wood from fungal infections, also increased NAG-1 mRNA and protein levels, along with induction of p53 in HCT-116 cells [206]. Similarly, silibinin, a flavonolignan extracted from milk thistle, exhibited antioxidant and antineoplastic activities, and increased NAG-1 mRNA and protein levels via the modulation of EGR-1 and p38 MAPK pathways, but not p53 or ATF3 in HT-29, a human colon carcinoma cell line [207].
In addition to colorectal cancer, several phytochemicals have been shown to inhibit prostate, lung and gastric cancer. 18α-Glycyrrhetinic acid, a bioactive triterpenoid in licorice, decreased the expression of pro-inflammatory cytokines, including HMGB1, IL-6, and IL-8, while increasing NAG-1 mRNA in a human prostate cancer cell line DU-145 [208]. Isochaihu-lactone also increased NAG-1 expression in LNCaP, a human prostate cancer cell line, with the induction of EGR-1, possibly by modulating JNK1/2 [209]. Treatment of A549 cells with isochaihu-lactone induced NAG-1 by activating EGR-1 on the NAG-1 promoter [210]. Human lung carcinoma cell lines A549 and NCI-H460 treated with Taiwanin A showed decreased cell viability with increased NAG-1 protein, and JNK inhibitors decreased NAG-1 mRNA levels [211]. Treatment of AGS, a human gastric adenocarcinoma cell line, with hispidulin, a naturally occurring flavone, also induced cell cycle arrest, apoptosis and NAG-1 expression, while COX-2 expression was down-regulated by the activation of ERK1/2 [212].
Overall, many phytochemicals with an anti-inflammatory activity increase NAG-1 expression in cancer cell lines by mechanisms involving EGR-1, p53, ATF3, C/EBPβ or kinase alterations. Two forms of NAG-1, i.e., pro- and mature form, exist in cells and both forms are likely to be up-regulated when cells are treated with phytochemicals. In addition to cancer, NAG-1 induction may help prevent obesity, which is a chronic inflammatory condition. NAG-1 induction results in the activation of the GFRAL receptor in the brain, causing a reduced appetite. Therefore, several NAG-1-inducing phytochemicals listed in Table 2 may prevent obesity and its associated diseases.

4.2. ROS-Inducing Natural Products as Anticancer Agents

Plant extracts have long been used to treat diverse ailments and diseases, and they are still popular remedies. Beginning in the 19th century, individual phytochemicals and microbial-derived compounds have rapidly developed into major sources of new drugs for chemotherapeutic applications [213,214,215]. Moreover, these natural products have also served as templates for—more potent derivatives/analogs such as the taxanes (paclitaxel) and the second-generation synthetic analogs. Several anticancer drugs, including natural product-derived anticancer agents such as curcumin, phenethyl-isothiocyanate (PEITC), benzyl-isothiocyanate (BITC), betulinic acid, and piper-longumine celastrol, and synthetic triterpenoids derived from glycherretinic acid and oleanolic acid are cell-specific ROS-inducing anticancer agents, and we have identified possible mechanisms of action for these ROS-inducing agents [216].
Table 2. Anti-inflammatory phytochemicals that increase NAG-1 expression.
Table 2. Anti-inflammatory phytochemicals that increase NAG-1 expression.
PhytochemicalCell LineDose (μM)Mechanism of ActionReference
2′-HydroxyflavanoneHCT-1165–40EGR-1[200]
6-GingerolHCT-11625–200PKCε, GSK-3β[217]
18α-Glycyrrhetinic acidDU-145100-[208]
ApigeninHCT-116, LoVo, SW480, HT-290.1–10PKCδ[201]
BerberineHCT-116, Caco-2, HepG21–100PKCε, GSK-3β, ERK1/2, EGR-1[218,219]
CAPEHCT-1161–25ATF3[203]
CapsaicinHCT-1161–100GSK3β, C/EBPβ, ATF3, PKCδ[204]
DamnacanthalHCT-116, LoVo1–100ERK, C/EBPβ[220]
Diallyl disulfideHCT-1164.6–23p53[221]
DIMHCT-11612.5–50ATF3[222]
Green tea (EGCG/ECG)HCT-1161–100ATF3, EGR-1[202,223,224]
GenisteinHCT-116, A54925–100p53[225,226]
FormononetinHCT-1166.25–400EGR-1[205]
HispidulinAGS6.25–100ERK1/2[212]
Indole-3-carbinolHCT-11625–100-[222]
IsochaihulactoneA549, LNCaP, GBM84011.25–80EGR-1, ERK1/2, JNK, DDIT3[209,210,227,228]
IsoliquiritigeninHCT-1162.5–160EGR-1[205]
Platycodon DU9377.5–15EGR-1[229]
PinosylvinHCT-11660p53[206]
Pseudolaric acid BHT-291–25EGR-1[230]
QuercetinHCT-116, Huh75–40EGR-1, p53[231]
ResveratrolHCT-116, A549, U2OS, S2-013, CD1810–100p53, RNA stability[232,233]
SilibininHT-2950–100EGR-1, p38 MAPK[207]
Taiwanin AA549, H4601.25–80JNK[211]
XanthorrhizolHCT-11625–100-[234]
The anticancer activities of curcumin, betulinic acid and methyl 2-cyano-3,12-dioxooleana-1,9-dien-28-oate (CDDO-Me, Bardoxolone methyl), a synthetic triterpenoid derived from oleanolic acid, exhibited several common features [235,236,237,238]. These compounds induced ROS, which is a critical response, and as co-treatment with antioxidants or catalase (to breakdown hydrogen peroxide) ameliorated most of the drug-induced inhibition of growth and migration/invasion and induction of apoptosis. In addition, these ROS inducers also downregulated expression of specificity protein (Sp) transcription factor (TFs) Sp1, Sp3 and Sp4, and several pro-oncogenic Sp-regulated genes, including survivin, bcl-2, receptor tyrosine kinases and VEGF. Thus, drug-dependent induction of ROS resulted in targeting Sp TFs, which was complemented by the studies showing that hydrogen peroxide, t-butyl hydroperoxide and other ROS/H2O2 inducers (e.g., arsenic trioxide and ascorbic acid) also decreased Sp TFs [235,239,240]. The functional and clinical significance of ROS-dependent downregulation of Sp TFs is supported by separate studies showing that knockdown of Sp1 or Sp3 and Sp4 (alone and in combination) decreased cancer cell growth, survival and migration/invasion, and tumor growth, mimicking the effects of ROS-inducing anticancer agents, in a mouse xenograft model [241,242].
Scott et al. [243] showed that microRNA-27a (miR-27a) suppressed the transcriptional repressor ZBTB10, which was previously reported to inhibit Sp1-dependent activation of gastric gene expression, possibly by competition with Sp1 for GC-rich Sp binding sites on the gastrin gene promoter [244]. Our results showed that miR-27a suppressed ZBTB10 in cancer cells and antisense miR-27a induced ZBTB10 expression with downregulation of Sp1, Sp3, Sp4 and pro-oncogenic Sp-regulated genes [245]. Subsequent studies showed that miR-27a also repressed ZBTB34 expression and miR-17-92 complex miRs block expression of ZBTB4 and both ZBTB34 and ZBTB4 overexpression downregulated Sp TFs [246]. The connection between ROS and the miR-ZBTB-Sp pathway was substantiated by several studies showing that ROS/ROS-inducing anticancer agents decreased miR expression and induced ZBTB4/10/34 which in turn downregulated Sp TFs and pro-oncogenic Sp-regulated genes [235,247,248,249,250].
O’Hagan et al. [251] treated colon cancer cells with hydrogen peroxide and observed genome-wide shifts of repressor complexes from non-GC-rich gene promoters, and this was accompanied by decreased expression of several genes, including cMyc. There is evidence that cMyc regulates the expression of multiple miRs, including miR-27a and miR-17-92, and we showed that ROS-inducing anticancer agents also decreased cMyc expression [248,249,250]. Subsequent studies showed that the effects of cMyc knockdown resemble those observed for ROS, resulting in decreased expression of miR-27a/miR-17-92, induction of ZBTB4/10/34 and downregulation of Sp TFs and pro-oncogenic Sp-regulated genes (Figure 2). Deviations from this pathway may be observed if one or more genes (miR/ZBTBs) are not expressed in a particular cancer cell. The identification of this ROS-mediated pathway will be important for the design of drug combination therapies where Sp TFs are a direct target, since multiple drugs induce this response, and some may be used interchangeably. ROS inducers also downregulate many Sp-regulated genes associated with drug resistance, and their use in combination drug therapies could counteract the development of drug resistance, which is commonly observed for many chemotherapies.

5. Identification of Biochemical Targets of Bioactive Compounds

Natural products have been used as medicines and remedies throughout human history. Diverse structures, high potency, and selectivity of natural products have led to the successful discovery of many drugs. Identifying molecular targets of natural products is pivotal in pharmacological research and drug discovery. This involves elucidation and characterization of the interactions between a bioactive molecule and its specific molecular target, which is commonly an enzyme or receptor. Four cases of our recent work are described here to illustrate how putative molecular targets have been initially identified. The intent is to share our lessons and experiences for biochemical target identification of bioactive compounds with serendipity and tools of mass spectrometry (MS) based proteomics, patch clamp, surface plasmon resonance, and bioassays.
The first is a serendipity case—identification of the C-glycosyl-flavone iso-orientin, isolated from corn silk, inhibiting glycogen synthase kinase-3β (GSK-3β) [252]. The result was a surprise because GSK-3β was the first kinase arbitrarily selected in our experiments to screen chemical isolates from corn silk after the isolates showed little activity against an acetyl cholinesterase. Overactive GSK-3β is responsible for hyperphosphorylation of tau protein. Isoorientin specifically inhibits GSK-3β via substrate competition. It effectively attenuates GSK-3β-catalyzed tau hyperphosphorylation and is neuroprotective against amyloid-induced neurotoxicity in human SH-SY5Y cells. It is known that hyperphosphorylation of tau proteins in neurons plays a pivotal role in the pathogenesis of Alzheimer’s disease (AD). Molecular, cellular and in vivo studies showed that isoorientin and its semi-synthetic analogs engage with GSK-3β, and they are relevant to neuroinflammation [253,254].
The second is use of MS-based quantitative proteomics to identify putative molecular targets [255]. Non-small cell lung cancer (NSCLC) is the most common type of lung cancer. The low efficacy in current chemotherapies calls for new alternatives to prevent or treat NSCLC [255]. 24-Methylenecyloartanyl ferulate (24-mCAF) isolated from rice bran oil is cytotoxic to NSCLC cells. The idea stemmed from the anti-inflammatory and anti-cancer activities of γ-oryzanol, a major component of rice bran oil [256]. An isobaric tag for relative and absolute quantitation-based quantitative proteomics analysis suggested that 24-mCAF inhibits cell proliferation and activates cell death and apoptosis in A549 cells. 24-mCAF up-regulates Myb binding protein 1A (MYBBP1A), a tumor suppressor. In vitro enzymatic assays confirmed that 24-mCAF inhibits the activity of AKT and Aurora B kinase, two Ser/Thr kinases involved in MYBBP1A regulation, and they may be important therapeutic targets in NSCLC.
In the third case, a potential neuron target was first noted by swift toxic action of putative compounds to fruit flies and extremely steep dose-mortality curves [257]. The monoterpenoids linalool, estragole, and methyl eugenol are the bioactive chemicals in many essential oils, such as basil oils. Basil oil and the monoterpenoids have been widely used for health benefits and pest insect control. However, the molecular target of those chemical constituents is not well understood. The γ-aminobutyric acid type A receptors (GABAAR) and nicotinic acetylcholine receptor (nAChR) are well-known primary molecular targets of synthetic insecticides. We, therefore, studied the electrophysiological effects of linalool, estragole, methyl eugenol, and citronellal on GABAAR and nAChR to further understand their versatility as traditional medicines and insecticides [258]. The results revealed that linalool inhibits both GABAAR and nAChR, which may explain its insecticidal activity. Linalool is a non-competitive inhibitor of GABAAR as evidenced by the negligible influences of linalool on the half-maximal effective values of GABA for the rat α1β3γ2L GABAAR. The half-maximal inhibitory concentration of linalool on the GABAAR was approximately 3.2 mM. As multiple monoterpenoids are present in the same essential oil, it is likely that linalool synergistically interacts with terpenoids, thus offering a possible explanation for its sedative and anticonvulsant effects as well as insecticidal activities [257,258].
The fourth case demonstrates how drug affinity responsive target stability (DARTS) and surface plasmon resonance (SPR) methods were used to identify biochemical targets of dihydromyricetin in 3T3-L1 adipocytes [259]. Dihydromyricetin, a flavanonol, is isolated from Ampelopsis grossedentata, an herb plant known for its weight-losing ability. Cell studies show that dihydromyricetin and epigallocatechin gallate effectively reduce intracellular oil droplet formation in 3T3-L1 cells. Although dihydromyricetin inhibits adipogenesis, the molecular target was not identified. DARTS and SPR experiments demonstrate the direct interactions of dihydromyricetin and EGCG with the 78-kDa glucose-regulated protein (GRP78), having a dissociation constant of 22 µM and 16 µM, respectively. GRP78 is essential for adipocyte differentiation and adipogenesis. Therefore, the results suggest a new understanding of dihydromyricetin in the modulation of obesity and its anti-obesity implications.

6. Nano-Delivery of Bioactive Compounds in Foods for Disease Prevention

Many food-derived bioactive compounds vary in their physical, chemical, and biological properties. They are present in a limited quantity in foods, and the complex structures of food matrices may limit their bioavailability. Extracting bioactive compounds with known bioactivities and incorporating them in processed food products may maximize their health benefits. However, there are several challenges. Many bioactive compounds are water-insoluble and are not distributed evenly in food matrices; many compounds are present as crystalline structures, leading to low bioavailability; and they can be degraded over time. The degradation is triggered or accelerated by factors, such as oxygen, UV, catalysts (e.g., multivalent cations), and heat [260]. As food products usually require a long shelf-life, the stability of bioactive compounds should be maintained during processing and storage. Lastly, some bioactive compounds are unstable in gastrointestinal conditions and are absorbed poorly, and some may present toxicity if their dose and safety are not adequately assessed [261].
Nanoparticles have unique properties in incorporating lipophilic bioactive compounds in foods. The ability to maintain the transparency of beverages is a unique feature of nanoparticles. However, the large surface area of nanoparticles may accelerate the degradation of encapsulated compounds during storage. Lipophilic compounds can be incorporated in biopolymers using various mechanisms such as molecular binding and manipulation of solubility. For example, anti-solvent precipitation is an encapsulation mechanism in which lipophilic bioactive compounds and a hydrophobic biopolymer are first dissolved in alcohol, followed by mixing in water to induce precipitation and encapsulation of the compound in biopolymer nanoparticles; alternatively, lipophilic bioactive compounds pre-dissolved in alcohol can be blended in a biopolymer solution to achieve encapsulation during mixing [261]. Lipophilic bioactive compounds can also be dissolved in lipid to prepare nanostructures in the forms of nano-emulsions, microemulsions, solid lipid nanoparticles, and nanostructured lipid carriers [261]. To improve the convenience of application and enable the use in solid foods, nanoparticles fabricated in liquid systems can be prepared in powdered form. Freeze drying is commonly used but may be expensive and less scalable than spray drying. The high temperature during spray drying may cause the degradation of bioactive compounds and the de-structuration of nanoparticles. Certain protectants may be used to minimize the impact of drying on chemical and physical properties of nanoparticles.
There are several factors to consider when developing nanotechnologies to incorporate bioactive compounds in foods. Both bioactive compounds such as curcumin, carotenoids, resveratrol, and polyunsaturated fatty acids and carrier materials such as proteins, polysaccharides, and lipids are ideally derived from natural products permitted for use in food products, although synthetic compounds with a structure identical to the natural ones, e.g., β-carotene, may be used as food additives [262]. Food-derived bioactive compounds are likely the starting point to develop food-based intervention strategies as these compounds may have fewer barriers in regulation, labeling, and consumer acceptance. The next step is to identify the dose needed for in vivo efficacy of each compound for disease prevention and the cost of the compounds. The availability and sustainability of bioactive compounds are critical for incorporation in foods. In food applications, purified bioactive compounds may be too costly and may not be needed, and extracts rich in these compounds may be more cost-effective. The next stage is to develop ingredients targeting specific food applications based on physicochemical characteristics, processing and storage conditions, and sensory properties. For liquid and semi-liquid products, encapsulation is likely to be needed to ensure even distribution and prevent phase separation during storage. For solid products, powdered ingredients, ideally encapsulation systems, are needed, which then require scalable and cost-effective encapsulation technologies and food-grade, sustainable encapsulation materials. The stability of bioactive compounds in foods during processing and storage is evaluated to ensure the quantity consumed and no harmful reaction products due to degradation. The encapsulated bioactive compounds, after incorporation in foods, are to be re-evaluated for toxicity and their effectiveness in preventing diseases, as these properties may be different from the encapsulation system alone [263]. Lastly, to determine the commercial feasibility, consumer acceptance in terms of sensory properties (appearance, texture, taste, and aroma) and labeling, regulatory compliance, and profitability of food products incorporated with bioactive compounds needs to be established.

7. Conclusions

Significant advances have been made in developing therapies to treat metabolic and chronic diseases triggered by inflammation and oxidative stress. However, several challenges and limitations remain, including drug resistance and side effects, which underscore the urgency of developing new and safe preventive approaches. Thus, natural bioactive compounds, including phytochemicals and various botanicals, emerged as promising alternatives to traditional pharmacological approaches. Researchers at the 20th Frontier Scientists Workshop discussed various natural products with antioxidant, anti-inflammatory, immune-modulatory, and anti-cancer activities. A major emphasis was on disease prevention via lifestyle intervention, including consumption of healthy foods that provide added benefits through the natural antioxidant and anti-inflammatory bioactive compounds they contain. Researchers at the Workshop concluded that continued efforts to identify and characterize new antioxidant and anti-inflammatory phytochemicals, their mechanisms of action, development of innovative technologies for incorporating them into our food system and increasing their bio-availability are warranted to effectively combat metabolic and chronic diseases that afflict modern society.

Author Contributions

All authors contributed equally to the manuscript preparation. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korean Academy of Science and Technology (KAST) and National R&D Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (NRF-2020R1A2B5B02002442).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mesika, R.; Reichmann, D. When safeguarding goes wrong: Impact of oxidative stress on protein homeostasis in health and neurodegenerative disorders. Adv. Protein Chem. Struct. Biol. 2019, 114, 221–264. [Google Scholar] [CrossRef]
  2. Singh, A.; Kukreti, R.; Saso, L.; Kukreti, S. Oxidative stress: A key modulator in neurodegenerative diseases. Molecules 2019, 24, 1583. [Google Scholar] [CrossRef] [Green Version]
  3. Karam, B.S.; Chavez-Moreno, A.; Koh, W.; Akar, J.G.; Akar, F.G. Oxidative stress and inflammation as central mediators of atrial fibrillation in obesity and diabetes. Cardiovasc. Diabetol. 2017, 16, 120. [Google Scholar] [CrossRef]
  4. Taniguchi, K.; Karin, M. NF-kappaB, inflammation, immunity and cancer: Coming of age. Nat. Rev. Immunol. 2018, 18, 309–324. [Google Scholar] [CrossRef] [PubMed]
  5. Todoric, J.; Karin, M. The fire within: Cell-Autonomous mechanisms in inflammation-driven cancer. Cancer Cell 2019, 35, 714–720. [Google Scholar] [CrossRef]
  6. Kim, G.H.; Kim, J.E.; Rhie, S.J.; Yoon, S. The role of oxidative stress in neurodegenerative diseases. Exp. Neurobiol. 2015, 24, 325–340. [Google Scholar] [CrossRef] [PubMed]
  7. Gupta, S.C.; Kunnumakkara, A.B.; Aggarwal, S.; Aggarwal, B.B. Inflammation, a double-edge sword for cancer and other age-related diseases. Front. Immunol. 2018, 9, 2160. [Google Scholar] [CrossRef]
  8. Albini, A.; DeCensi, A.; Cavalli, F.; Costa, A. Cancer prevention and interception: A new era for chemopreventive approaches. Clin. Cancer Res. 2016, 22, 4322–4327. [Google Scholar] [CrossRef] [Green Version]
  9. Ballon-Landa, E.; Parsons, J.K. Nutrition, physical activity, and lifestyle factors in prostate cancer prevention. Curr. Opin. Urol. 2018, 28, 55–61. [Google Scholar] [CrossRef]
  10. Albini, A.; Bassani, B.; Baci, D.; Dallaglio, K.; Gallazzi, M.; Corradino, P.; Bruno, A.; Noonan, D.M. Nutraceuticals and “repurposed” drugs of phytochemical origin in prevention and interception of chronic degenerative diseases and cancer. Curr. Med. Chem. 2019, 26, 973–987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Sporn, M.B.; Liby, K.T.; Yore, M.M.; Fu, L.; Lopchuk, J.M.; Gribble, G.W. New synthetic triterpenoids: Potent agents for prevention and treatment of tissue injury caused by inflammatory and oxidative stress. J. Nat. Prod. 2011, 74, 537–545. [Google Scholar] [CrossRef]
  12. Yang, C.S.; Luo, P.; Zeng, Z.; Wang, H.; Malafa, M.; Suh, N. Vitamin E and cancer prevention: Studies with different forms of tocopherols and tocotrienols. Mol. Carcinog. 2020, 59, 365–389. [Google Scholar] [CrossRef]
  13. Suh, N.; Wang, Y.; Honda, T.; Gribble, G.W.; Dmitrovsky, E.; Hickey, W.F.; Maue, R.A.; Place, A.E.; Porter, D.M.; Spinella, M.J.; et al. A novel synthetic oleanane triterpenoid, 2-cyano-3,12-dioxoolean-1,9-dien-28-oic acid, with potent differentiating, antiproliferative, and anti-inflammatory activity. Cancer Res. 1999, 59, 336–341. [Google Scholar] [PubMed]
  14. Dinkova-Kostova, A.T.; Liby, K.T.; Stephenson, K.K.; Holtzclaw, W.D.; Gao, X.; Suh, N.; Williams, C.; Risingsong, R.; Honda, T.; Gribble, G.W.; et al. Extremely potent triterpenoid inducers of the phase 2 response: Correlations of protection against oxidant and inflammatory stress. Proc. Natl. Acad. Sci. USA 2005, 102, 4584–4589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Honda, T.; Gribble, G.W.; Suh, N.; Finlay, H.J.; Rounds, B.V.; Bore, L.; Favaloro, F.G., Jr.; Wang, Y.; Sporn, M.B. Novel synthetic oleanane and ursane triterpenoids with various enone functionalities in ring A as inhibitors of nitric oxide production in mouse macrophages. J. Med. Chem. 2000, 43, 1866–1877. [Google Scholar] [CrossRef] [PubMed]
  16. Liby, K.T.; Sporn, M.B. Synthetic oleanane triterpenoids: Multifunctional drugs with a broad range of applications for prevention and treatment of chronic disease. Pharm. Rev. 2012, 64, 972–1003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Jiang, Q. Natural forms of vitamin E: Metabolism, antioxidant, and anti-inflammatory activities and their role in disease prevention and therapy. Free Radic. Biol. Med. 2014, 72, 76–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Ju, J.; Picinich, S.C.; Yang, Z.; Zhao, Y.; Suh, N.; Kong, A.N.; Yang, C.S. Cancer-preventive activities of tocopherols and tocotrienols. Carcinogenesis 2010, 31, 533–542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Wang, X.; Quinn, P.J. Vitamin E and its function in membranes. Prog. Lipid Res. 1999, 38, 309–336. [Google Scholar] [CrossRef]
  20. Mustacich, D.J.; Bruno, R.S.; Traber, M.G. Vitamin E. Vitam. Horm. 2007, 76, 1–21. [Google Scholar] [CrossRef]
  21. Niki, E.; Traber, M.G. A history of vitamin E. Ann. Nutr. Metab. 2012, 61, 207–212. [Google Scholar] [CrossRef]
  22. Traber, M.G.; Stevens, J.F. Vitamins C and E: Beneficial effects from a mechanistic perspective. Free Radic. Biol. Med. 2011, 51, 1000–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Traber, M.G. Vitamin E regulatory mechanisms. Annu. Rev. Nutr. 2007, 27, 347–362. [Google Scholar] [CrossRef]
  24. Yang, C.S.; Chen, J.X.; Wang, H.; Lim, J. Lessons learned from cancer prevention studies with nutrients and non-nutritive dietary constituents. Mol. Nutr. Food Res. 2016, 60, 1239–1250. [Google Scholar] [CrossRef] [PubMed]
  25. Brigelius-Flohe, R.; Traber, M.G. Vitamin E: Function and metabolism. FASEB J. 1999, 13, 1145–1155. [Google Scholar] [CrossRef] [PubMed]
  26. Hensley, K.; Benaksas, E.J.; Bolli, R.; Comp, P.; Grammas, P.; Hamdheydari, L.; Mou, S.; Pye, Q.N.; Stoddard, M.F.; Wallis, G.; et al. New perspectives on vitamin E: Gamma-tocopherol and carboxyelthylhydroxychroman metabolites in biology and medicine. Free Radic. Biol. Med. 2004, 36, 1–15. [Google Scholar] [CrossRef] [PubMed]
  27. Kamal-Eldin, A.; Appelqvist, L.A. The chemistry and antioxidant properties of tocopherols and tocotrienols. Lipids 1996, 31, 671–701. [Google Scholar] [CrossRef]
  28. Lippman, S.M.; Klein, E.A.; Goodman, P.J.; Lucia, M.S.; Thompson, I.M.; Ford, L.G.; Parnes, H.L.; Minasian, L.M.; Gaziano, J.M.; Hartline, J.A.; et al. Effect of selenium and vitamin E on risk of prostate cancer and other cancers: The Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA 2009, 301, 39–51. [Google Scholar] [CrossRef] [Green Version]
  29. Klein, E.A.; Thompson, I.M., Jr.; Tangen, C.M.; Crowley, J.J.; Lucia, M.S.; Goodman, P.J.; Minasian, L.M.; Ford, L.G.; Parnes, H.L.; Gaziano, J.M.; et al. Vitamin E and the risk of prostate cancer: The selenium and Vitamin E cancer prevention trial (SELECT). JAMA 2011, 306, 1549–1556. [Google Scholar] [CrossRef] [PubMed]
  30. Chen, J.X.; Liu, A.; Lee, M.J.; Wang, H.; Yu, S.; Chi, E.; Reuhl, K.; Suh, N.; Yang, C.S. Delta- and gamma-tocopherols inhibit phIP/DSS-induced colon carcinogenesis by protection against early cellular and DNA damages. Mol. Carcinog. 2017, 56, 172–183. [Google Scholar] [CrossRef] [Green Version]
  31. Smolarek, A.K.; So, J.Y.; Thomas, P.E.; Lee, H.J.; Paul, S.; Dombrowski, A.; Wang, C.X.; Saw, C.L.; Khor, T.O.; Kong, A.N.; et al. Dietary tocopherols inhibit cell proliferation, regulate expression of ERalpha, PPARgamma, and Nrf2, and decrease serum inflammatory markers during the development of mammary hyperplasia. Mol. Carcinog. 2013, 52, 514–525. [Google Scholar] [CrossRef] [Green Version]
  32. Smolarek, A.K.; So, J.Y.; Burgess, B.; Kong, A.N.; Reuhl, K.; Lin, Y.; Shih, W.J.; Li, G.; Lee, M.J.; Chen, Y.K.; et al. Dietary administration of delta- and gamma-tocopherol inhibits tumorigenesis in the animal model of estrogen receptor-positive, but not HER-2 breast cancer. Cancer Prev. Res. Phila 2012, 5, 1310–1320. [Google Scholar] [CrossRef] [Green Version]
  33. Guan, F.; Li, G.; Liu, A.B.; Lee, M.J.; Yang, Z.; Chen, Y.K.; Lin, Y.; Shih, W.; Yang, C.S. Delta- and gamma-tocopherols, but not alpha-tocopherol, inhibit colon carcinogenesis in azoxymethane-treated F344 rats. Cancer Prev. Res. Phila 2012, 5, 644–654. [Google Scholar] [CrossRef] [Green Version]
  34. Bak, M.J.; Das Gupta, S.; Wahler, J.; Lee, H.J.; Li, X.; Lee, M.J.; Yang, C.S.; Suh, N. Inhibitory effects of gamma- and delta-Tocopherols on Estrogen-Stimulated breast cancer In Vitro and In Vivo. Cancer Prev. Res. Phila 2017, 10, 188–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Das Gupta, S.; Patel, M.; Wahler, J.; Bak, M.J.; Wall, B.; Lee, M.J.; Lin, Y.; Shih, W.J.; Cai, L.; Yang, C.S.; et al. Differential gene regulation and tumor-inhibitory activities of alpha-, delta-, and gamma-tocopherols in estrogen-mediated mammary carcinogenesis. Cancer Prev. Res. Phila 2017, 10, 694–703. [Google Scholar] [CrossRef] [Green Version]
  36. Das Gupta, S.; Sae-tan, S.; Wahler, J.; So, J.Y.; Bak, M.J.; Cheng, L.C.; Lee, M.J.; Lin, Y.; Shih, W.J.; Shull, J.D.; et al. Dietary gamma-tocopherol-rich mixture inhibits estrogen-induced mammary tumorigenesis by modulating estrogen metabolism, antioxidant response, and PPARgamma. Cancer Prev. Res. Phila 2015, 8, 807–816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Das Gupta, S.; So, J.Y.; Wall, B.; Wahler, J.; Smolarek, A.K.; Sae-Tan, S.; Soewono, K.Y.; Yu, H.; Lee, M.J.; Thomas, P.E.; et al. Tocopherols inhibit oxidative and nitrosative stress in estrogen-induced early mammary hyperplasia in ACI rats. Mol. Carcinog. 2015, 54, 916–925. [Google Scholar] [CrossRef] [Green Version]
  38. Christen, S.; Woodall, A.A.; Shigenaga, M.K.; Southwell-Keely, P.T.; Duncan, M.W.; Ames, B.N. Gamma-tocopherol traps mutagenic electrophiles such as NO(X) and complements alpha-tocopherol: Physiological implications. Proc. Natl. Acad. Sci. USA 1997, 94, 3217–3222. [Google Scholar] [CrossRef] [Green Version]
  39. Hoglen, N.C.; Waller, S.C.; Sipes, I.G.; Liebler, D.C. Reactions of peroxynitrite with gamma-tocopherol. Chem. Res. Toxicol. 1997, 10, 401–407. [Google Scholar] [CrossRef]
  40. Jiang, Q.; Lykkesfeldt, J.; Shigenaga, M.K.; Shigeno, E.T.; Christen, S.; Ames, B.N. Gamma-tocopherol supplementation inhibits protein nitration and ascorbate oxidation in rats with inflammation. Free Radic. Biol. Med. 2002, 33, 1534–1542. [Google Scholar] [CrossRef]
  41. Ogawa, Y.; Saito, Y.; Nishio, K.; Yoshida, Y.; Ashida, H.; Niki, E. Gamma-tocopheryl quinone, not alpha-tocopheryl quinone, induces adaptive response through up-regulation of cellular glutathione and cysteine availability via activation of ATF4. Free Radic. Res. 2008, 42, 674–687. [Google Scholar] [CrossRef]
  42. Jiang, Q.; Ames, B.N. Gamma-tocopherol, but not alpha-tocopherol, decreases proinflammatory eicosanoids and inflammation damage in rats. FASEB J. 2003, 17, 816–822. [Google Scholar] [CrossRef]
  43. Jiang, Q.; Elson-Schwab, I.; Courtemanche, C.; Ames, B.N. Gamma-tocopherol and its major metabolite, in contrast to alpha-tocopherol, inhibit cyclooxygenase activity in macrophages and epithelial cells. Proc. Natl. Acad. Sci. USA 2000, 97, 11494–11499. [Google Scholar] [CrossRef] [Green Version]
  44. Reiter, E.; Jiang, Q.; Christen, S. Anti-inflammatory properties of alpha- and gamma-tocopherol. Mol. Asp. Med. 2007, 28, 668–691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. O’Leary, K.A.; de Pascual-Tereasa, S.; Needs, P.W.; Bao, Y.P.; O’Brien, N.M.; Williamson, G. Effect of flavonoids and vitamin E on cyclooxygenase-2 (COX-2) transcription. Mutat. Res. 2004, 551, 245–254. [Google Scholar] [CrossRef] [PubMed]
  46. Kalupahana, N.S.; Moustaid-Moussa, N.; Claycombe, K.J. Immunity as a link between obesity and insulin resistance. Mol. Asp. Med. 2012, 33, 26–34. [Google Scholar] [CrossRef]
  47. Salvestrini, V.; Sell, C.; Lorenzini, A. Obesity may accelerate the aging process. Front. Endocrinol. Lausanne 2019, 10, 266. [Google Scholar] [CrossRef] [PubMed]
  48. Siriwardhana, N.; Kalupahana, N.S.; Cekanova, M.; LeMieux, M.; Greer, B.; Moustaid-Moussa, N. Modulation of adipose tissue inflammation by bioactive food compounds. J. Nutr. Biochem. 2013, 24, 613–623. [Google Scholar] [CrossRef] [PubMed]
  49. Jayarathne, S.; Stull, A.J.; Park, O.H.; Kim, J.H.; Thompson, L.; Moustaid-Moussa, N. Protective effects of anthocyanins in obesity-associated inflammation and changes in gut microbiome. Mol. Nutr. Food Res. 2019, 63, e1900149. [Google Scholar] [CrossRef]
  50. Thangthaeng, N.; Poulose, S.M.; Gomes, S.M.; Miller, M.G.; Bielinski, D.F.; Shukitt-Hale, B. Tart cherry supplementation improves working memory, hippocampal inflammation, and autophagy in aged rats. Age Dordr. 2016, 38, 393–404. [Google Scholar] [CrossRef] [Green Version]
  51. Kalupahana, N.S.; Claycombe, K.J.; Moustaid-Moussa, N. (n-3) Fatty acids alleviate adipose tissue inflammation and insulin resistance: Mechanistic insights. Adv. Nutr. 2011, 2, 304–316. [Google Scholar] [CrossRef] [Green Version]
  52. Jayarathne, S.; Stull, A.J.; Miranda, A.; Scoggin, S.; Claycombe-Larson, K.; Kim, J.H.; Moustaid-Moussa, N. Tart cherry reduces inflammation in adipose tissue of zucker fatty rats and cultured 3T3-L1 adipocytes. Nutrients 2018, 10, 1576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kirakosyan, A.; Seymour, E.M.; Llanes, D.E.U.; Kaufman, P.B.; Bolling, S.F. Chemical profile and antioxidant capacities of tart cherry products. Food Chem. 2009, 115, 20–25. [Google Scholar] [CrossRef]
  54. Seymour, E.M.; Lewis, S.K.; Urcuyo-Llanes, D.E.; Tanone, I.I.; Kirakosyan, A.; Kaufman, P.B.; Bolling, S.F. Regular tart cherry intake alters abdominal adiposity, adipose gene transcription, and inflammation in obesity-prone rats fed a high fat diet. J. Med. Food 2009, 12, 935–942. [Google Scholar] [CrossRef]
  55. Kim, B.; Lee, S.G.; Park, Y.K.; Ku, C.S.; Pham, T.X.; Wegner, C.J.; Yang, Y.; Koo, S.I.; Chun, O.K.; Lee, J.Y. Blueberry, blackberry, and blackcurrant differentially affect plasma lipids and pro-inflammatory markers in diet-induced obesity mice. Nutr. Res. Pract. 2016, 10, 494–500. [Google Scholar] [CrossRef]
  56. Kolehmainen, M.; Mykkanen, O.; Kirjavainen, P.V.; Leppanen, T.; Moilanen, E.; Adriaens, M.; Laaksonen, D.E.; Hallikainen, M.; Puupponen-Pimia, R.; Pulkkinen, L.; et al. Bilberries reduce low-grade inflammation in individuals with features of metabolic syndrome. Mol. Nutr. Food Res. 2012, 56, 1501–1510. [Google Scholar] [CrossRef] [Green Version]
  57. McCall, D.O.; McGartland, C.P.; McKinley, M.C.; Sharpe, P.; McCance, D.R.; Young, I.S.; Woodside, J.V. The effect of increased dietary fruit and vegetable consumption on endothelial activation, inflammation and oxidative stress in hypertensive volunteers. Nutr. Metab. Cardiovasc. Dis. 2011, 21, 658–664. [Google Scholar] [CrossRef]
  58. Meydani, M. Nutrition interventions in aging and age-associated disease. Ann. N. Y. Acad. Sci. 2001, 928, 226–235. [Google Scholar] [CrossRef] [PubMed]
  59. Jayarathne, S.; Ramalingam, L.; Edwards, H.; Vanapalli, S.A.; Moustaid-Moussa, N. Tart cherry increases lifespan in caenorhabditis elegans by altering metabolic signaling pathways. Nutrients 2020, 12, 1482. [Google Scholar] [CrossRef] [PubMed]
  60. Liguori, I.; Russo, G.; Curcio, F.; Bulli, G.; Aran, L.; Della-Morte, D.; Gargiulo, G.; Testa, G.; Cacciatore, F.; Bonaduce, D. Oxidative stress, aging, and diseases. Clin. Interv. Aging 2018, 13, 757. [Google Scholar] [CrossRef] [Green Version]
  61. Shen, P.; Yue, Y.; Zheng, J.; Park, Y. Caenorhabditis elegans: A convenient in vivo model for assessing the impact of food bioactive compounds on obesity, aging, and Alzheimer’s disease. Annu. Rev. Food Sci. Technol. 2018, 9, 1–22. [Google Scholar] [CrossRef] [PubMed]
  62. Olsen, A.; Vantipalli, M.C.; Lithgow, G.J. Using Caenorhabditis elegans as a model for aging and age-related diseases. Ann. N. Y. Acad. Sci. 2006, 1067, 120–128. [Google Scholar] [CrossRef] [PubMed]
  63. Sueishi, Y.; Nii, R.; Kakizaki, N. Resveratrol analogues like piceatannol are potent antioxidants as quantitatively demonstrated through the high scavenging ability against reactive oxygen species and methyl radical. Bioorg. Med. Chem. Lett. 2017, 27, 5203–5206. [Google Scholar] [CrossRef] [PubMed]
  64. Kiliç, I.; Yeşiloğlu, Y. Spectroscopic studies on the antioxidant activity of p-coumaric acid. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2013, 115, 719–724. [Google Scholar] [CrossRef]
  65. Thygesen, L.; Thulin, J.; Mortensen, A.; Skibsted, L.H.; Molgaard, P. Antioxidant activity of cichoric acid and alkamides from Echinacea purpurea, alone and in combination. Food Chem. 2007, 101, 74–81. [Google Scholar] [CrossRef]
  66. Shen, P.; Yue, Y.; Sun, Q.; Kasireddy, N.; Kim, K.H.; Park, Y. Piceatannol extends the lifespan of Caenorhabditis elegans via DAF-16. Biofactors 2017, 43, 379–387. [Google Scholar] [CrossRef]
  67. Peng, Y.; Sun, Q.; Gao, R.; Park, Y. AAK-2 and SKN-1 are involved in chicoric-acid-induced lifespan extension in Caenorhabditis elegans. J. Agric. Food Chem. 2019, 67, 9178–9186. [Google Scholar] [CrossRef]
  68. Yue, Y.; Shen, P.; Xu, Y.; Park, Y. p-Coumaric acid improves oxidative and osmosis stress responses in Caenorhabditis elegans. J. Sci. Food Agric. 2019, 99, 1190–1197. [Google Scholar] [CrossRef]
  69. Farina, F.; Lambert, E.; Commeau, L.; Lejeune, F.-X.; Roudier, N.; Fonte, C.; Parker, J.A.; Boddaert, J.; Verny, M.; Baulieu, E.-E. The stress response factor daf-16/FOXO is required for multiple compound families to prolong the function of neurons with Huntington’s disease. Sci. Rep. 2017, 7, 1–15. [Google Scholar] [CrossRef] [Green Version]
  70. Lee, J.; Kwon, G.; Park, J.; Kim, J.-K.; Lim, Y.-H. Brief communication: SIR-2.1-dependent lifespan extension of Caenorhabditis elegans by oxyresveratrol and resveratrol. Exp. Biol. Med. 2016, 241, 1757–1763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Xiong, L.-G.; Huang, J.-A.; Li, J.; Yu, P.-H.; Xiong, Z.; Zhang, J.-W.; Gong, Y.-S.; Liu, Z.-H.; Chen, J.-H. Black tea increased survival of Caenorhabditis elegans under stress. J. Agric. Food Chem. 2014, 62, 11163–11169. [Google Scholar] [CrossRef]
  72. Aparecida Paiva, F.; de Freitas Bonomo, L.; Ferreira Boasquivis, P.; Borges Raposo de Paula, I.T.; Guerra, J.F.d.C.; Mendes Leal, W.; Silva, M.E.; Pedrosa, M.L.; Oliveira, R.d.P. Carqueja (Baccharis trimera) protects against oxidative stress and β-amyloid-induced toxicity in Caenorhabditis elegans. Oxidative Med. Cell. Longev. 2015, 2015, 740162. [Google Scholar] [CrossRef] [Green Version]
  73. Grünz, G.; Haas, K.; Soukup, S.; Klingenspor, M.; Kulling, S.E.; Daniel, H.; Spanier, B. Structural features and bioavailability of four flavonoids and their implications for lifespan-extending and antioxidant actions in C. elegans. Mech. Ageing Dev. 2012, 133, 1–10. [Google Scholar] [CrossRef]
  74. Shen, P.; Yue, Y.; Park, Y. A Living model for obesity and aging research: Caenorhabditis elegans. Crit. Rev. Food Sci. Nutr. 2017, 58, 741–754. [Google Scholar] [CrossRef] [PubMed]
  75. Bansal, A.; Zhu, L.J.; Yen, K.; Tissenbaum, H.A. Uncoupling lifespan and healthspan in Caenorhabditis elegans longevity mutants. Proc. Natl. Acad. Sci. USA 2015, 112, E277–E286. [Google Scholar] [CrossRef] [Green Version]
  76. Harrell, C.R.; Jankovic, M.G.; Fellabaum, C.; Volarevic, A.; Djonov, V.; Arsenijevic, A.; Volarevic, V. Molecular mechanisms responsible for anti-inflammatory and immunosuppressive effects of mesenchymal stem cell-derived factors. Adv. Exp. Med. Biol. 2019, 1084, 187–206. [Google Scholar] [CrossRef] [PubMed]
  77. Li, Y.; Wu, Q.; Wang, Y.; Li, L.; Bu, H.; Bao, J. Senescence of mesenchymal stem cells (Review). Int. J. Mol. Med. 2017, 39, 775–782. [Google Scholar] [CrossRef] [Green Version]
  78. Yu, K.R.; Kang, K.S. Aging-related genes in mesenchymal stem cells: A mini-review. Gerontology 2013, 59, 557–563. [Google Scholar] [CrossRef] [PubMed]
  79. Kim, S.; Yoon, Y.M.; Han, Y.S.; Lee, J.H.; Hur, J.; Lee, S.H. Administration of Cripto in GRP78 overexpressed human MSCs enhances stem cell viability and angiogenesis during human MSC transplantation therapy. Cell Prolif. 2018, 51, e12463. [Google Scholar] [CrossRef] [PubMed]
  80. Lee, J.H.; Han, Y.S.; Lee, S.H. Potentiation of biological effects of mesenchymal stem cells in ischemic conditions by melatonin via upregulation of cellular prion protein expression. J. Pineal Res. 2017, 62, e12385. [Google Scholar] [CrossRef] [PubMed]
  81. Lee, H.J.; Jung, Y.H.; Choi, G.E.; Kim, J.S.; Chae, C.W.; Lim, J.R.; Kim, S.Y.; Lee, J.E.; Park, M.C.; Yoon, J.H.; et al. O-cyclic phytosphingosine-1-phosphate stimulates HIF1alpha-dependent glycolytic reprogramming to enhance the therapeutic potential of mesenchymal stem cells. Cell Death Dis. 2019, 10, 590. [Google Scholar] [CrossRef] [PubMed]
  82. Forni, C.; Facchiano, F.; Bartoli, M.; Pieretti, S.; Facchiano, A.; D’Arcangelo, D.; Norelli, S.; Valle, G.; Nisini, R.; Beninati, S.; et al. Beneficial role of phytochemicals on oxidative stress and age-related diseases. Biomed. Res. Int. 2019, 2019, 8748253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Saud, B.; Malla, R.; Shrestha, K. A review on the effect of plant extract on mesenchymal stem cell proliferation and differentiation. Stem Cells Int. 2019, 2019, 7513404. [Google Scholar] [CrossRef] [PubMed]
  84. Lee, H.J.; Jung, Y.H.; Oh, J.Y.; Choi, G.E.; Chae, C.W.; Kim, J.S.; Lim, J.R.; Kim, S.Y.; Lee, S.J.; Seong, J.K.; et al. BICD1 mediates HIF1α nuclear translocation in mesenchymal stem cells during hypoxia adaptation. Cell Death Differ. 2019, 26, 1716–1734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Van der Bliek, A.M.; Sedensky, M.M.; Morgan, P.G. Cell biology of the mitochondrion. Genetics 2017, 207, 843–871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Ko, S.H.; Choi, G.E.; Oh, J.Y.; Lee, H.J.; Kim, J.S.; Chae, C.W.; Choi, D.; Han, H.J. Author correction: Succinate promotes stem cell migration through the GPR91-dependent regulation of DRP1-mediated mitochondrial fission. Sci. Rep. 2018, 8, 13326. [Google Scholar] [CrossRef] [Green Version]
  87. Lee, H.J.; Jung, Y.H.; Choi, G.E.; Ko, S.H.; Lee, S.J.; Lee, S.H.; Han, H.J. BNIP3 induction by hypoxia stimulates FASN-dependent free fatty acid production enhancing therapeutic potential of umbilical cord blood-derived human mesenchymal stem cells. Redox Biol. 2017, 13, 426–443. [Google Scholar] [CrossRef]
  88. Liu, S.P.; Shibu, M.A.; Tsai, F.J.; Hsu, Y.M.; Tsai, C.H.; Chung, J.G.; Yang, J.S.; Tang, C.H.; Wang, S.; Li, Q.; et al. Tetramethylpyrazine reverses high-glucose induced hypoxic effects by negatively regulating HIF-1alpha induced BNIP3 expression to ameliorate H9c2 cardiomyoblast apoptosis. Nutr. Metab. 2020, 17, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Grabacka, M.M.; Gawin, M.; Pierzchalska, M. Phytochemical modulators of mitochondria: The search for chemopreventive agents and supportive therapeutics. Pharmaceuticals 2014, 7, 913–942. [Google Scholar] [CrossRef] [Green Version]
  90. Sharifi, S.; Zununi Vahed, S.; Ahmadian, E.; Maleki Dizaj, S.; Abedi, A.; Hosseiniyan Khatibi, S.M.; Samiei, M. Stem cell therapy: Curcumin does the trick. Phytother. Res. 2019, 33, 2927–2937. [Google Scholar] [CrossRef] [Green Version]
  91. Mohar, D.S.; Malik, S. The Sirtuin system: The holy grail of resveratrol? J. Clin. Exp. Cardiol. 2012, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Jung, Y.H.; Lee, H.J.; Kim, J.S.; Lee, S.J.; Han, H.J. EphB2 signaling-mediated Sirt3 expression reduces MSC senescence by maintaining mitochondrial ROS homeostasis. Free Radic. Biol. Med. 2017, 110, 368–380. [Google Scholar] [CrossRef] [PubMed]
  93. Oh, J.Y.; Choi, G.E.; Lee, H.J.; Jung, Y.H.; Chae, C.W.; Kim, J.S.; Lee, C.K.; Han, H.J. 17beta-Estradiol protects mesenchymal stem cells against high glucose-induced mitochondrial oxidants production via Nrf2/Sirt3/MnSOD signaling. Free Radic. Biol. Med. 2019, 130, 328–342. [Google Scholar] [CrossRef] [PubMed]
  94. Lee, J.H.; Yun, C.W.; Hur, J.; Lee, S.H. Fucoidan rescues p-cresol-induced cellular senescence in mesenchymal stem cells via FAK-Akt-TWIST axis. Mar. Drugs 2018, 16, 121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Lee, J.H.; Ryu, J.M.; Han, Y.S.; Zia, M.F.; Kwon, H.Y.; Noh, H.; Han, H.J.; Lee, S.H. Fucoidan improves bioactivity and vasculogenic potential of mesenchymal stem cells in murine hind limb ischemia associated with chronic kidney disease. J. Mol. Cell. Cardiol. 2016, 97, 169–179. [Google Scholar] [CrossRef] [PubMed]
  96. Han, Y.S.; Lee, J.H.; Jung, J.S.; Noh, H.; Baek, M.J.; Ryu, J.M.; Yoon, Y.M.; Han, H.J.; Lee, S.H. Fucoidan protects mesenchymal stem cells against oxidative stress and enhances vascular regeneration in a murine hindlimb ischemia model. Int. J. Cardiol. 2015, 198, 187–195. [Google Scholar] [CrossRef]
  97. Carey, M.A.; Germolec, D.R.; Langenbach, R.; Zeldin, D.C. Cyclooxygenase enzymes in allergic inflammation and asthma. Prostaglandins Leukot. Essent. Fat. Acids 2003, 69, 157–162. [Google Scholar] [CrossRef]
  98. Peebles, R.S., Jr.; Hashimoto, K.; Morrow, J.D.; Dworski, R.; Collins, R.D.; Hashimoto, Y.; Christman, J.W.; Kang, K.H.; Jarzecka, K.; Furlong, J.; et al. Selective cyclooxygenase-1 and -2 inhibitors each increase allergic inflammation and airway hyperresponsiveness in mice. Am. J. Respir. Crit. Care Med. 2002, 165, 1154–1160. [Google Scholar] [CrossRef]
  99. Halushka, M.K.; Walker, L.P.; Halushka, P.V. Genetic variation in cyclooxygenase 1: Effects on response to aspirin. Clin. Pharm. 2003, 73, 122–130. [Google Scholar] [CrossRef]
  100. Szczeklik, W.; Sanak, M.; Szczeklik, A. Functional effects and gender association of COX-2 gene polymorphism G-765C in bronchial asthma. J. Allergy Clin. Immunol. 2004, 114, 248–253. [Google Scholar] [CrossRef]
  101. Lee, C.R.; Bottone, F.G., Jr.; Krahn, J.M.; Li, L.; Mohrenweiser, H.W.; Cook, M.E.; Petrovich, R.M.; Bell, D.A.; Eling, T.E.; Zeldin, D.C. Identification and functional characterization of polymorphisms in human cyclooxygenase-1 (PTGS1). Pharm. Genom. 2007, 17, 145–160. [Google Scholar] [CrossRef] [Green Version]
  102. Kabashima, K.; Murata, T.; Tanaka, H.; Matsuoka, T.; Sakata, D.; Yoshida, N.; Katagiri, K.; Kinashi, T.; Tanaka, T.; Miyasaka, M.; et al. Thromboxane A2 modulates interaction of dendritic cells and T cells and regulates acquired immunity. Nat. Immunol. 2003, 4, 694–701. [Google Scholar] [CrossRef]
  103. Katamura, K.; Shintaku, N.; Yamauchi, Y.; Fukui, T.; Ohshima, Y.; Mayumi, M.; Furusho, K. Prostaglandin E2 at priming of naive CD4+ T cells inhibits acquisition of ability to produce IFN-gamma and IL-2, but not IL-4 and IL-5. J. Immunol. 1995, 155, 4604–4612. [Google Scholar]
  104. Zhou, W.; Blackwell, T.S.; Goleniewska, K.; O’Neal, J.F.; Fitzgerald, G.A.; Lucitt, M.; Breyer, R.M.; Peebles, R.S., Jr. Prostaglandin I2 analogs inhibit Th1 and Th2 effector cytokine production by CD4 T cells. J. Leukoc. Biol. 2007, 81, 809–817. [Google Scholar] [CrossRef] [PubMed]
  105. Carey, M.A.; Germolec, D.R.; Bradbury, J.A.; Gooch, R.A.; Moorman, M.P.; Flake, G.P.; Langenbach, R.; Zeldin, D.C. Accentuated T helper type 2 airway response after allergen challenge in cyclooxygenase-1-/- but not cyclooxygenase-2-/- mice. Am. J. Respir. Crit. Care Med. 2003, 167, 1509–1515. [Google Scholar] [CrossRef]
  106. Li, H.; Bradbury, J.A.; Dackor, R.T.; Edin, M.L.; Graves, J.P.; DeGraff, L.M.; Wang, P.M.; Bortner, C.D.; Maruoka, S.; Lih, F.B.; et al. Cyclooxygenase-2 regulates Th17 cell differentiation during allergic lung inflammation. Am. J. Respir. Crit. Care Med. 2011, 184, 37–49. [Google Scholar] [CrossRef]
  107. Durrant, D.M.; Metzger, D.W. Emerging roles of T helper subsets in the pathogenesis of asthma. Immunol. Investig. 2010, 39, 526–549. [Google Scholar] [CrossRef]
  108. Li, H.; Edin, M.L.; Bradbury, J.A.; Graves, J.P.; DeGraff, L.M.; Gruzdev, A.; Cheng, J.; Dackor, R.T.; Wang, P.M.; Bortner, C.D.; et al. Cyclooxygenase-2 inhibits T helper cell type 9 differentiation during allergic lung inflammation via down-regulation of IL-17RB. Am. J. Respir. Crit. Care Med. 2013, 187, 812–822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Mahdi, J.G.; Mahdi, A.J.; Mahdi, A.J.; Bowen, I.D. The historical analysis of aspirin discovery, its relation to the willow tree and antiproliferative and anticancer potential. Cell Prolif. 2006, 39, 147–155. [Google Scholar] [CrossRef] [PubMed]
  110. Vane, J.R.; Botting, R.M. The mechanism of action of aspirin. Thromb. Res. 2003, 110, 255–258. [Google Scholar] [CrossRef]
  111. Attiq, A.; Jalil, J.; Husain, K.; Ahmad, W. Raging the war against inflammation with natural products. Front. Pharmacol. 2018, 9, 976. [Google Scholar] [CrossRef]
  112. Bagga, D.; Wang, L.; Farias-Eisner, R.; Glaspy, J.A.; Reddy, S.T. Differential effects of prostaglandin derived from omega-6 and omega-3 polyunsaturated fatty acids on COX-2 expression and IL-6 secretion. Proc. Natl. Acad. Sci. USA 2003, 100, 1751–1756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Li, H.; Edin, M.L.; Gruzdev, A.; Cheng, J.; Bradbury, J.A.; Graves, J.P.; DeGraff, L.M.; Zeldin, D.C. Regulation of T helper cell subsets by cyclooxygenases and their metabolites. Prostaglandins Other Lipid Mediat. 2013, 104–105, 74–83. [Google Scholar] [CrossRef] [PubMed]
  114. Askew, R.L.; Capo-Lugo, C.E.; Naidech, A.; Prabhakaran, S. Differential effects of time to initiation of therapy on disability and quality of life in patients with mild and moderate to severe ischemic stroke. Arch. Phys. Med. Rehabil. 2020, 101, 1515–1522. [Google Scholar] [CrossRef]
  115. Khoshnam, S.E.; Winlow, W.; Farzaneh, M.; Farbood, Y.; Moghaddam, H.F. Pathogenic mechanisms following ischemic stroke. Neurol. Sci. 2017, 38, 1167–1186. [Google Scholar] [CrossRef] [PubMed]
  116. Siesjo, B.K. Pathophysiology and treatment of focal cerebral ischemia. Part I: Pathophysiology. J. Neurosurg. 1992, 77, 169–184. [Google Scholar] [CrossRef]
  117. Song, D.; Xu, J.; Du, T.; Yan, E.; Hertz, L.; Walz, W.; Peng, L. Inhibition of brain swelling after ischemia-reperfusion by beta-adrenergic antagonists: Correlation with increased K+ and decreased Ca2+ concentrations in extracellular fluid. Biomed. Res. Int 2014, 2014, 873590. [Google Scholar] [CrossRef] [Green Version]
  118. Chen, H.; Yoshioka, H.; Kim, G.S.; Jung, J.E.; Okami, N.; Sakata, H.; Maier, C.M.; Narasimhan, P.; Goeders, C.E.; Chan, P.H. Oxidative stress in ischemic brain damage: Mechanisms of cell death and potential molecular targets for neuroprotection. Antioxid. Redox Signal. 2011, 14, 1505–1517. [Google Scholar] [CrossRef] [Green Version]
  119. Casas, A.I.; Geuss, E.; Kleikers, P.W.M.; Mencl, S.; Herrmann, A.M.; Buendia, I.; Egea, J.; Meuth, S.G.; Lopez, M.G.; Kleinschnitz, C.; et al. NOX4-dependent neuronal autotoxicity and BBB breakdown explain the superior sensitivity of the brain to ischemic damage. Proc. Natl. Acad. Sci. USA 2017, 114, 12315–12320. [Google Scholar] [CrossRef] [Green Version]
  120. Huang, Y.; Chen, S.; Luo, Y.; Han, Z. Crosstalk between Inflammation and the BBB in Stroke. Curr. Neuropharmacol. 2020, 18, 1227–1236. [Google Scholar] [CrossRef]
  121. Fu, Y.; Liu, Q.; Anrather, J.; Shi, F.D. Immune interventions in stroke. Nat. Rev. Neurol. 2015, 11, 524–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Keaney, J.; Campbell, M. The dynamic blood-brain barrier. FEBS J. 2015, 282, 4067–4079. [Google Scholar] [CrossRef] [PubMed]
  123. Suzuki, R.; Yamaguchi, T.; Kirino, T.; Orzi, F.; Klatzo, I. The effects of 5-minute ischemia in Mongolian gerbils: I. Blood-brain barrier, cerebral blood flow, and local cerebral glucose utilization changes. Acta Neuropathol. 1983, 60, 207–216. [Google Scholar] [CrossRef] [PubMed]
  124. Pillai, D.R.; Dittmar, M.S.; Baldaranov, D.; Heidemann, R.M.; Henning, E.C.; Schuierer, G.; Bogdahn, U.; Schlachetzki, F. Cerebral ischemia-reperfusion injury in rats—a 3 T MRI study on biphasic blood-brain barrier opening and the dynamics of edema formation. J. Cereb. Blood Flow Metab. 2009, 29, 1846–1855. [Google Scholar] [CrossRef]
  125. Knowland, D.; Arac, A.; Sekiguchi, K.J.; Hsu, M.; Lutz, S.E.; Perrino, J.; Steinberg, G.K.; Barres, B.A.; Nimmerjahn, A.; Agalliu, D. Stepwise recruitment of transcellular and paracellular pathways underlies blood-brain barrier breakdown in stroke. Neuron 2014, 82, 603–617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Shi, Y.; Zhang, L.; Pu, H.; Mao, L.; Hu, X.; Jiang, X.; Xu, N.; Stetler, R.A.; Zhang, F.; Liu, X.; et al. Rapid endothelial cytoskeletal reorganization enables early blood-brain barrier disruption and long-term ischaemic reperfusion brain injury. Nat. Commun. 2016, 7, 10523. [Google Scholar] [CrossRef]
  127. Song, J.; Wu, C.; Korpos, E.; Zhang, X.; Agrawal, S.M.; Wang, Y.; Faber, C.; Schafers, M.; Korner, H.; Opdenakker, G.; et al. Focal MMP-2 and MMP-9 activity at the blood-brain barrier promotes chemokine-induced leukocyte migration. Cell Rep. 2015, 10, 1040–1054. [Google Scholar] [CrossRef] [Green Version]
  128. Romanic, A.M.; White, R.F.; Arleth, A.J.; Ohlstein, E.H.; Barone, F.C. Matrix metalloproteinase expression increases after cerebral focal ischemia in rats: Inhibition of matrix metalloproteinase-9 reduces infarct size. Stroke 1998, 29, 1020–1030. [Google Scholar] [CrossRef] [Green Version]
  129. Candelario-Jalil, E.; Yang, Y.; Rosenberg, G.A. Diverse roles of matrix metalloproteinases and tissue inhibitors of metalloproteinases in neuroinflammation and cerebral ischemia. Neuroscience 2009, 158, 983–994. [Google Scholar] [CrossRef] [Green Version]
  130. Malemud, C.J. Matrix metalloproteinases (MMPs) in health and disease: An overview. Front. Biosci. 2006, 11, 1696–1701. [Google Scholar] [CrossRef]
  131. Yang, Y.; Estrada, E.Y.; Thompson, J.F.; Liu, W.; Rosenberg, G.A. Matrix metalloproteinase-mediated disruption of tight junction proteins in cerebral vessels is reversed by synthetic matrix metalloproteinase inhibitor in focal ischemia in rat. J. Cereb. Blood Flow Metab. 2007, 27, 697–709. [Google Scholar] [CrossRef] [PubMed]
  132. Maier, C.M.; Hsieh, L.; Crandall, T.; Narasimhan, P.; Chan, P.H. Evaluating therapeutic targets for reperfusion-related brain hemorrhage. Ann. Neurol. 2006, 59, 929–938. [Google Scholar] [CrossRef]
  133. Simard, M.; Arcuino, G.; Takano, T.; Liu, Q.S.; Nedergaard, M. Signaling at the gliovascular interface. J. Neurosci. 2003, 23, 9254–9262. [Google Scholar] [CrossRef] [PubMed]
  134. Solenov, E.; Watanabe, H.; Manley, G.T.; Verkman, A.S. Sevenfold-reduced osmotic water permeability in primary astrocyte cultures from AQP-4-deficient mice, measured by a fluorescence quenching method. Am. J. Physiol. Cell Physiol. 2004, 286, C426–C432. [Google Scholar] [CrossRef]
  135. Manley, G.T.; Fujimura, M.; Ma, T.; Noshita, N.; Filiz, F.; Bollen, A.W.; Chan, P.; Verkman, A.S. Aquaporin-4 deletion in mice reduces brain edema after acute water intoxication and ischemic stroke. Nat. Med. 2000, 6, 159–163. [Google Scholar] [CrossRef] [PubMed]
  136. Hyun, S.W.; Jung, Y.S. Hypoxia induces FoxO3a-mediated dysfunction of blood-brain barrier. Biochem. Biophys. Res. Commun. 2014, 450, 1638–1642. [Google Scholar] [CrossRef] [PubMed]
  137. Zhang, Z.G.; Zhang, L.; Jiang, Q.; Zhang, R.; Davies, K.; Powers, C.; Bruggen, N.; Chopp, M. VEGF enhances angiogenesis and promotes blood-brain barrier leakage in the ischemic brain. J. Clin. Investig. 2000, 106, 829–838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Wu, L.; Ye, Z.; Pan, Y.; Li, X.; Fu, X.; Zhang, B.; Li, Y.; Lin, W.; Li, X.; Gao, Q. Vascular endothelial growth factor aggravates cerebral ischemia and reperfusion-induced blood-brain-barrier disruption through regulating LOC102640519/HOXC13/ZO-1 signaling. Exp. Cell Res. 2018, 369, 275–283. [Google Scholar] [CrossRef]
  139. Ji, B.; Zhou, F.; Han, L.; Yang, J.; Fan, H.; Li, S.; Li, J.; Zhang, X.; Wang, X.; Chen, X.; et al. Sodium tanshinone IIA sulfonate enhances effectiveness Rt-PA treatment in acute ischemic stroke patients associated with ameliorating blood-brain barrier damage. Transl. Stroke Res. 2017, 8, 334–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Lee, K.; Lee, J.S.; Jang, H.J.; Kim, S.M.; Chang, M.S.; Park, S.H.; Kim, K.S.; Bae, J.; Park, J.W.; Lee, B.; et al. Chlorogenic acid ameliorates brain damage and edema by inhibiting matrix metalloproteinase-2 and 9 in a rat model of focal cerebral ischemia. Eur. J. Pharm. 2012, 689, 89–95. [Google Scholar] [CrossRef]
  141. Zhang, X.; Fan, Z.; Jin, T. Crocin protects against cerebral- ischemia-induced damage in aged rats through maintaining the integrity of blood-brain barrier. Restor. Neurol. Neurosci. 2017, 35, 65–75. [Google Scholar] [CrossRef] [PubMed]
  142. Ma, Y.; Li, L.; Kong, L.; Zhu, Z.; Zhang, W.; Song, J.; Chang, J.; Du, G. Pinocembrin protects blood-brain barrier function and expands the therapeutic time window for tissue-type plasminogen activator treatment in a rat thromboembolic stroke model. Biomed. Res. Int. 2018, 2018, 8943210. [Google Scholar] [CrossRef]
  143. Zhang, W.; Song, J.K.; Zhang, X.; Zhou, Q.M.; He, G.R.; Xu, X.N.; Rong, Y.; Zhou, W.X.; Du, G.H. Salvianolic acid A attenuates ischemia reperfusion induced rat brain damage by protecting the blood brain barrier through MMP-9 inhibition and anti-inflammation. Chin. J. Nat. Med. 2018, 16, 184–193. [Google Scholar] [CrossRef]
  144. Li, S.; Bian, L.; Fu, X.; Ai, Q.; Sui, Y.; Zhang, A.; Gao, H.; Zhong, L.; Lu, D. Gastrodin pretreatment alleviates rat brain injury caused by cerebral ischemic-reperfusion. Brain Res. 2019, 1712, 207–216. [Google Scholar] [CrossRef]
  145. Chang, C.Y.; Chen, J.Y.; Wu, M.H.; Hu, M.L. Therapeutic treatment with vitamin C reduces focal cerebral ischemia-induced brain infarction in rats by attenuating disruptions of blood brain barrier and cerebral neuronal apoptosis. Free Radic. Biol. Med. 2020, 155, 29–36. [Google Scholar] [CrossRef] [PubMed]
  146. Liu, C.D.; Liu, N.N.; Zhang, S.; Ma, G.D.; Yang, H.G.; Kong, L.L.; Du, G.H. Salvianolic acid A prevented cerebrovascular endothelial injury caused by acute ischemic stroke through inhibiting the Src signaling pathway. Acta Pharm. Sin. 2020, 42, 370–381. [Google Scholar] [CrossRef] [PubMed]
  147. Li, M.; Ma, R.N.; Li, L.H.; Qu, Y.Z.; Gao, G.D. Astragaloside IV reduces cerebral edema post-ischemia/reperfusion correlating the suppression of MMP-9 and AQP4. Eur. J. Pharm. 2013, 715, 189–195. [Google Scholar] [CrossRef]
  148. Wang, Y.; Wu, Y.; Liang, C.; Tan, R.; Tan, L.; Tan, R. Pharmacodynamic effect of ellagic acid on ameliorating cerebral ischemia/reperfusion injury. Pharmacology 2019, 104, 320–331. [Google Scholar] [CrossRef]
  149. Nag, S.; Venugopalan, R.; Stewart, D.J. Increased caveolin-1 expression precedes decreased expression of occludin and claudin-5 during blood-brain barrier breakdown. Acta Neuropathol. 2007, 114, 459–469. [Google Scholar] [CrossRef]
  150. Lee, B.K.; Hyun, S.W.; Jung, Y.S. Yuzu and hesperidin ameliorate blood-brain barrier disruption during hypoxia via antioxidant activity. Antioxid. Basel 2020, 9, 843. [Google Scholar] [CrossRef]
  151. Liu, J.; Chen, L.; Zhang, X.; Pan, L.; Jiang, L. The protective effects of juglanin in cerebral ischemia reduce blood-brain barrier permeability via inhibition of VEGF/VEGFR2 signaling. Drug Des. Dev. Ther. 2020, 14, 3165–3175. [Google Scholar] [CrossRef]
  152. Chen, H.; Guan, B.; Chen, X.; Chen, X.; Li, C.; Qiu, J.; Yang, D.; Liu, K.J.; Qi, S.; Shen, J. Baicalin attenuates blood-brain barrier disruption and hemorrhagic transformation and improves neurological outcome in ischemic stroke rats with delayed t-PA treatment: Involvement of ONOO(-)-MMP-9 pathway. Transl. Stroke Res. 2018, 9, 515–529. [Google Scholar] [CrossRef]
  153. Li, W.; Suwanwela, N.C.; Patumraj, S. Curcumin prevents reperfusion injury following ischemic stroke in rats via inhibition of NFkappaB, ICAM-1, MMP-9 and caspase-3 expression. Mol. Med. Rep. 2017, 16, 4710–4720. [Google Scholar] [CrossRef] [Green Version]
  154. Li, J.; Liu, Y.; Zhang, X.; Chen, R.; Zhang, L.; Xue, J.; Gao, X. Dl-3-N-Butylphthalide alleviates the blood-brain barrier permeability of focal cerebral ischemia reperfusion in mice. Neuroscience 2019, 413, 99–107. [Google Scholar] [CrossRef] [PubMed]
  155. Chen, W.; Guo, Y.; Yang, W.; Zheng, P.; Zeng, J.; Tong, W. Protective effect of ginsenoside Rb1 on integrity of blood-brain barrier following cerebral ischemia. Exp. Brain Res. 2015, 233, 2823–2831. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, X.; Liu, X.; Hu, G.; Zhang, G.; Zhao, G.; Shi, M. Ginsenoside Rd attenuates blood-brain barrier damage by suppressing proteasome-mediated signaling after transient forebrain ischemia. Neuroreport 2020, 31, 466–472. [Google Scholar] [CrossRef] [PubMed]
  157. Liu, M.B.; Wang, W.; Gao, J.M.; Li, F.; Shi, J.S.; Gong, Q.H. Icariside II attenuates cerebral ischemia/reperfusion-induced blood-brain barrier dysfunction in rats via regulating the balance of MMP9/TIMP1. Acta Pharm. Sin. 2020, 41, 1547–1556. [Google Scholar] [CrossRef]
  158. Jang, J.W.; Lee, J.K.; Lee, M.C.; Piao, M.S.; Kim, S.H.; Kim, H.S. Melatonin reduced the elevated matrix metalloproteinase-9 level in a rat photothrombotic stroke model. J. Neurol. Sci. 2012, 323, 221–227. [Google Scholar] [CrossRef]
  159. Jin, Z.; Ke, J.; Guo, P.; Wang, Y.; Wu, H. Quercetin improves blood-brain barrier dysfunction in rats with cerebral ischemia reperfusion via Wnt signaling pathway. Am. J. Transl. Res. 2019, 11, 4683–4695. [Google Scholar]
  160. Wei, H.; Wang, S.; Zhen, L.; Yang, Q.; Wu, Z.; Lei, X.; Lv, J.; Xiong, L.; Xue, R. Resveratrol attenuates the blood-brain barrier dysfunction by regulation of the MMP-9/TIMP-1 balance after cerebral ischemia reperfusion in rats. J. Mol. Neurosci. 2015, 55, 872–879. [Google Scholar] [CrossRef] [PubMed]
  161. Jang, J.W.; Lee, J.K.; Hur, H.; Kim, T.W.; Joo, S.P.; Piao, M.S. Rutin improves functional outcome via reducing the elevated matrix metalloproteinase-9 level in a photothrombotic focal ischemic model of rats. J. Neurol. Sci. 2014, 339, 75–80. [Google Scholar] [CrossRef]
  162. Mondal, N.K.; Behera, J.; Kelly, K.E.; George, A.K.; Tyagi, P.K.; Tyagi, N. Tetrahydrocurcumin epigenetically mitigates mitochondrial dysfunction in brain vasculature during ischemic stroke. Neurochem. Int. 2019, 122, 120–138. [Google Scholar] [CrossRef] [PubMed]
  163. Gong, P.; Zhang, Z.; Zou, Y.; Tian, Q.; Han, S.; Xu, Z.; Liao, J.; Gao, L.; Chen, Q.; Li, M. Tetramethylpyrazine attenuates blood-brain barrier disruption in ischemia/reperfusion injury through the JAK/STAT signaling pathway. Eur. J. Pharm. 2019, 854, 289–297. [Google Scholar] [CrossRef] [PubMed]
  164. Kodani, S.D.; Hammock, B.D. The 2014 Bernard, B. Brodie award lecture-epoxide hydrolases: Drug metabolism to therapeutics for chronic pain. Drug Metab. Dispos. 2015, 43, 788–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Wagner, K.M.; McReynolds, C.B.; Schmidt, W.K.; Hammock, B.D. Soluble epoxide hydrolase as a therapeutic target for pain, inflammatory and neurodegenerative diseases. Pharmacol. Ther. 2017, 180, 62–76. [Google Scholar] [CrossRef]
  166. Inceoglu, B.; Bettaieb, A.; Trindade da Silva, C.A.; Lee, K.S.; Haj, F.G.; Hammock, B.D. Endoplasmic reticulum stress in the peripheral nervous system is a significant driver of neuropathic pain. Proc. Natl. Acad. Sci. USA. 2015, 112, 9082–9087. [Google Scholar] [CrossRef] [Green Version]
  167. Cervenka, L.; Huskova, Z.; Kopkan, L.; Kikerlova, S.; Sedlakova, L.; Vanourkova, Z.; Alanova, P.; Kolar, F.; Hammock, B.D.; Hwang, S.H.; et al. Two pharmacological epoxyeicosatrienoic acid-enhancing therapies are effectively antihypertensive and reduce the severity of ischemic arrhythmias in rats with angiotensin II-dependent hypertension. J. Hypertens. 2018, 36, 1326–1341. [Google Scholar] [CrossRef]
  168. Chábová, Č.V.; Kujal, P.; Škaroupková, P.; Varňourková, Z.; Vacková, Š.; Husková, Z.; Kikerlová, S.; Sadowski, J.; Kompanowska-Jezierska, E.; Baranowska, I.; et al. Combined Inhibition of soluble epoxide hydrolase and renin-angiotensin system exhibits superior renoprotection to renin-angiotensin system blockade in 5/6 nephrectomized ren-2 transgenic hypertensive rats with established chronic kidney disease. Kidney Blood Press. Res. 2018, 43, 329–349. [Google Scholar] [CrossRef]
  169. Zarriello, S.; Tuazon, J.P.; Corey, S.; Schimmel, S.; Rajani, M.; Gorsky, A.; Incontri, D.; Hammock, B.D.; Borlongan, C.V. Humble beginnings with big goals: Small molecule soluble epoxide hydrolase inhibitors for treating CNS disorders. Prog. Neurobiol. 2019, 172, 23–39. [Google Scholar] [CrossRef]
  170. Swardfager, W.; Hennebelle, M.; Yu, D.; Hammock, B.D.; Levitt, A.J.; Hashimoto, K.; Taha, A.Y. Metabolic/inflammatory/vascular comorbidity in psychiatric disorders; soluble epoxide hydrolase (sEH) as a possible new target. Neurosci. Biobehav. Rev. 2018, 87, 56–66. [Google Scholar] [CrossRef] [Green Version]
  171. Tu, R.; Armstrong, J.; Lee, K.S.S.; Hammock, B.D.; Sapirstein, A.; Koehler, R.C. Soluble epoxide hydrolase inhibition decreases reperfusion injury after focal cerebral ischemia. Sci. Rep. 2018, 8, 5279. [Google Scholar] [CrossRef]
  172. Ren, Q.; Ma, M.; Yang, J.; Nonaka, R.; Yamaguchi, A.; Ishikawa, K.I.; Kobayashi, K.; Murayama, S.; Hwang, S.H.; Saiki, S.; et al. Soluble epoxide hydrolase plays a key role in the pathogenesis of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2018, 115, E5815–E5823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Ma, M.; Ren, Q.; Yang, J.; Zhang, K.; Xiong, Z.; Ishima, T.; Pu, Y.; Hwang, S.H.; Toyoshima, M.; Iwayama, Y.; et al. Key role of soluble epoxide hydrolase in the neurodevelopmental disorders of offspring after maternal immune activation. Proc. Natl. Acad. Sci. USA 2019, 116, 7083–7088. [Google Scholar] [CrossRef] [Green Version]
  174. Liang, Z.; Zhang, B.; Xu, M.; Morisseau, C.; Hwang, S.H.; Hammock, B.D.; Li, Q.X. 1-Trifluoromethoxyphenyl-3-(1-propionylpiperidin-4-yl) urea, a selective and potent dual inhibitor of soluble epoxide hydrolase and p38 kinase intervenes in alzheimer’s signaling in human nerve cells. ACS Chem. Neurosci. 2019, 10, 4018–4030. [Google Scholar] [CrossRef]
  175. Wang, W.; Yang, J.; Edin, M.L.; Wang, Y.; Luo, Y.; Wan, D.; Yang, H.; Song, C.Q.; Xue, W.; Sanidad, K.Z.; et al. Targeted metabolomics identifies the cytochrome P450 monooxygenase eicosanoid pathway as a novel therapeutic target of colon tumorigenesis. Cancer Res. 2019, 79, 1822–1830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Panigrahy, D.; Gartung, A.; Yang, J.; Yang, H.; Gilligan, M.M.; Sulciner, M.L.; Bhasin, S.S.; Bielenberg, D.R.; Chang, J.; Schmidt, B.A.; et al. Preoperative stimulation of resolution and inflammation blockade eradicates micrometastases. J. Clin. Investig. 2019, 129, 2964–2979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Guedes, A.; Galuppo, L.; Hood, D.; Hwang, S.H.; Morisseau, C.; Hammock, B.D. Soluble epoxide hydrolase activity and pharmacologic inhibition in horses with chronic severe laminitis. Equine Vet. J. 2017, 49, 345–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. McReynolds, C.B.; Hwang, S.H.; Yang, J.; Wan, D.; Wagner, K.; Morisseau, C.; Li, D.; Schmidt, W.K.; Hammock, B.D. Pharmaceutical effects of inhibiting the soluble epoxide hydrolase in canine osteoarthritis. Front. Pharmacol. 2019, 10, 533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Hammock, B.D.; McReynolds, C.B.; Wagner, K.; Buckpitt, A.; Cortes-Puch, I.; Croston, G.; Lee, K.S.S.; Yang, J.; Schmidt, W.K.; Hwang, S.H. Movement to the clinic of soluble epoxide hydrolase inhibitor EC5026 as an analgesic for neuropathic pain and for use as a nonaddictive opioid alternative. J. Med. Chem. 2021, 64, 1856–1872. [Google Scholar] [CrossRef]
  180. McReynolds, C.B.; Cortes-Puch, I.; Ravindran, R.; Khan, I.H.; Hammock, B.G.; Shih, P.-a.B.; Hammock, B.D.; Yang, J. Plasma Linoleate Diols Are Potential Biomarkers for Severe COVID-19 Infections. Front. Physiol. 2021, 12, 663869. [Google Scholar] [CrossRef]
  181. Ferrara, M.; De Gennaro, L. How much sleep do we need? Sleep Med. Rev. 2001, 5, 155–179. [Google Scholar] [CrossRef] [PubMed]
  182. Kim, G.H.; Kim, Y.; Yoon, S.; Kim, S.J.; Yi, S.S. Sleep-inducing effect of Passiflora incarnata L. extract by single and repeated oral administration in rodent animals. Food Sci. Nutr. 2020, 8, 557–566. [Google Scholar] [CrossRef] [PubMed]
  183. Roth, T. Insomnia: Definition, prevalence, etiology, and consequences. J. Clin. Sleep Med. 2007, 3, S7–S10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Shan, L.L.; Guo, H.; Song, N.N.; Jia, Z.P.; Hu, X.T.; Huang, J.F.; Ding, Y.Q.; Richter-Levin, G.; Zhou, Q.X.; Xu, L. Light exposure before learning improves memory consolidation at night. Sci. Rep. 2015, 5, 15578. [Google Scholar] [CrossRef]
  185. Lucassen, E.A.; Zhao, X.; Rother, K.I.; Mattingly, M.S.; Courville, A.B.; de Jonge, L.; Csako, G.; Cizza, G. Sleep Extension Study Group. Evening chronotype is associated with changes in eating behavior, more sleep apnea, and increased stress hormones in short sleeping obese individuals. PLoS ONE 2013, 8, e56519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Khan, S.; Duan, P.; Yao, L.; Hou, H. Shiftwork-Mediated disruptions of circadian rhythms and sleep homeostasis cause serious health problems. Int. J. Genom. 2018, 2018, 8576890. [Google Scholar] [CrossRef] [PubMed]
  187. Potter, G.D.; Skene, D.J.; Arendt, J.; Cade, J.E.; Grant, P.J.; Hardie, L.J. Circadian rhythm and sleep disruption: Causes, metabolic consequences, and countermeasures. Endocr. Rev. 2016, 37, 584–608. [Google Scholar] [CrossRef] [Green Version]
  188. Brites, D.; Fernandes, A. Neuroinflammation and depression: Microglia activation, extracellular microvesicles and microRNA dysregulation. Front. Cell Neurosci. 2015, 9, 476. [Google Scholar] [CrossRef] [Green Version]
  189. Leproult, R.; Van Cauter, E. Role of sleep and sleep loss in hormonal release and metabolism. Endocr. Dev. 2010, 17, 11–21. [Google Scholar] [CrossRef] [Green Version]
  190. Kim, G.H.; Lim, K.; Yang, H.S.; Lee, J.K.; Kim, Y.; Park, S.K.; Kim, S.H.; Park, S.; Kim, T.H.; Moon, J.S.; et al. Improvement in neurogenesis and memory function by administration of Passiflora incarnata L. extract applied to sleep disorder in rodent models. J. Chem. Neuroanat. 2019, 98, 27–40. [Google Scholar] [CrossRef]
  191. Miroddi, M.; Calapai, G.; Navarra, M.; Minciullo, P.L.; Gangemi, S. Passiflora incarnata L.: Ethnopharmacology, clinical application, safety and evaluation of clinical trials. J. Ethnopharmacol. 2013, 150, 791–804. [Google Scholar] [CrossRef]
  192. He, M.; Min, J.-W.; Kong, W.-L.; He, X.-H.; Li, J.-X.; Peng, B.-W. A review on the pharmacological effects of vitexin and isovitexin. Fitoterapia 2016, 115, 74–85. [Google Scholar] [CrossRef] [PubMed]
  193. Malar, D.S.; Suryanarayanan, V.; Prasanth, M.I.; Singh, S.K.; Balamurugan, K.; Devi, K.P. Vitexin inhibits Abeta25-35 induced toxicity in Neuro-2a cells by augmenting Nrf-2/HO-1 dependent antioxidant pathway and regulating lipid homeostasis by the activation of LXR-alpha. Toxicol. Vitr. 2018, 50, 160–171. [Google Scholar] [CrossRef] [PubMed]
  194. Soulimani, R.; Younos, C.; Jarmouni, S.; Bousta, D.; Misslin, R.; Mortier, F. Behavioural effects of Passiflora incarnata L. and its indole alkaloid and flavonoid derivatives and maltol in the mouse. J. Ethnopharmacol. 1997, 57, 11–20. [Google Scholar] [CrossRef]
  195. Choi, J.S.; Islam, M.N.; Ali, M.Y.; Kim, E.J.; Kim, Y.M.; Jung, H.A. Effects of C-glycosylation on anti-diabetic, anti-Alzheimer’s disease and anti-inflammatory potential of apigenin. Food Chem. Toxicol. 2014, 64, 27–33. [Google Scholar] [CrossRef] [PubMed]
  196. Abbasi, B.; Kimiagar, M.; Sadeghniiat, K.; Shirazi, M.M.; Hedayati, M.; Rashidkhani, B. The effect of magnesium supplementation on primary insomnia in elderly: A double-blind placebo-controlled clinical trial. J. Res. Med. Sci. 2012, 17, 1161–1169. [Google Scholar]
  197. Akhondzadeh, S.; Naghavi, H.R.; Vazirian, M.; Shayeganpour, A.; Rashidi, H.; Khani, M. Passionflower in the treatment of generalized anxiety: A pilot double-blind randomized controlled trial with oxazepam. J. Clin. Pharm. Ther. 2001, 26, 363–367. [Google Scholar] [CrossRef] [PubMed]
  198. Shal, B.; Ding, W.; Ali, H.; Kim, Y.S.; Khan, S. Anti-neuroinflammatory potential of natural products in attenuation of alzheimer’s disease. Front. Pharm. 2018, 9, 548. [Google Scholar] [CrossRef]
  199. Lasry, A.; Zinger, A.; Ben-Neriah, Y. Inflammatory networks underlying colorectal cancer. Nat. Immunol. 2016, 17, 230–240. [Google Scholar] [CrossRef]
  200. Shin, S.Y.; Kim, J.H.; Lee, J.H.; Lim, Y.; Lee, Y.H. 2’-Hydroxyflavanone induces apoptosis through Egr-1 involving expression of Bax, p21, and NAG-1 in colon cancer cells. Mol. Nutr. Food Res. 2012, 56, 761–774. [Google Scholar] [CrossRef]
  201. Zhong, Y.; Krisanapun, C.; Lee, S.-H.; Nualsanit, T.; Sams, C.; Peungvicha, P.; Baek, S.J. Molecular targets of apigenin in colorectal cancer cells: Involvement of p21, NAG-1 and p53. Eur. J. Cancer 2010, 46, 3365–3374. [Google Scholar] [CrossRef] [Green Version]
  202. Kang, S.U.; Lee, B.-S.; Lee, S.-H.; Baek, S.J.; Shin, Y.S.; Kim, C.-H. Expression of NSAID-activated gene-1 by EGCG in head and neck cancer: Involvement of ATM-dependent p53 expression. J. Nutr. Biochem. 2013, 24, 986–999. [Google Scholar] [CrossRef]
  203. Park, M.-H.; Chung, C.; Lee, S.H.; Baek, S.J.; Kim, J.S. Caffeic acid phenethyl ester induces the expression of NAG-1 via activating transcription factor 3. J. Life Sci. 2018, 28, 37–42. [Google Scholar] [CrossRef]
  204. Lee, S.H.; Krisanapun, C.; Baek, S.J. NSAID-activated gene-1 as a molecular target for capsaicin-induced apoptosis through a novel molecular mechanism involving GSK3beta, C/EBPbeta and ATF3. Carcinogenesis 2010, 31, 719–728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Auyeung, K.K.; Ko, J.K. Novel herbal flavonoids promote apoptosis but differentially induce cell cycle arrest in human colon cancer cell. Investig. New Drugs 2010, 28, 1–13. [Google Scholar] [CrossRef] [PubMed]
  206. Park, E.-J.; Chung, H.-J.; Park, H.J.; Kim, G.D.; Ahn, Y.-H.; Lee, S.K. Suppression of Src/ERK and GSK-3/β-catenin signaling by pinosylvin inhibits the growth of human colorectal cancer cells. Food Chem. Toxicol. 2013, 55, 424–433. [Google Scholar] [CrossRef] [PubMed]
  207. Woo, S.M.; Min, K.J.; Kim, S.; Park, J.W.; Kim, D.E.; Chun, K.S.; Kim, Y.H.; Lee, T.J.; Kim, S.H.; Choi, Y.H.; et al. Silibinin induces apoptosis of HT29 colon carcinoma cells through early growth response-1 (EGR-1)-mediated non-steroidal anti-inflammatory drug-activated gene-1 (NAG-1) up-regulation. Chem. Biol. Interact. 2014, 211, 36–43. [Google Scholar] [CrossRef] [PubMed]
  208. Shetty, A.V.; Thirugnanam, S.; Dakshinamoorthy, G.; Samykutty, A.; Zheng, G.; Chen, A.; Bosland, M.C.; Kajdacsy-Balla, A.; Gnanasekar, M. 18α-glycyrrhetinic acid targets prostate cancer cells by down-regulating inflammation-related genes. Int. J. Oncol. 2011, 39, 635–640. [Google Scholar] [CrossRef] [Green Version]
  209. Chiu, S.C.; Wang, M.J.; Yang, H.H.; Chen, S.P.; Huang, S.Y.; Chen, Y.L.; Lin, S.Z.; Harn, H.J.; Pang, C.Y. Activation of NAG-1 via JNK signaling revealed an isochaihulactone-triggered cell death in human LNCaP prostate cancer cells. BMC Cancer 2011, 11, 146. [Google Scholar] [CrossRef] [Green Version]
  210. Yu, Y.L.; Su, K.J.; Chen, C.J.; Wei, C.W.; Lin, C.J.; Yiang, G.T.; Lin, S.Z.; Harn, H.J.; Chen, Y.L. Synergistic anti-tumor activity of isochaihulactone and paclitaxel on human lung cancer cells. J. Cell Physiol. 2012, 227, 213–222. [Google Scholar] [CrossRef]
  211. Harn, H.-J.; Chuang, H.-M.; Chang, L.-F.; Huang, A.; Hsieh, S.-T.; Lin, S.-Z.; Chou, C.-W.; Kuo, Y.-H.; Chiou, T.-W. Taiwanin A targets non-steroidal anti-inflammatory drug-activated gene-1 in human lung carcinoma. Fitoterapia 2014, 99, 227–235. [Google Scholar] [CrossRef]
  212. Yu, C.Y.; Su, K.Y.; Lee, P.L.; Jhan, J.Y.; Tsao, P.H.; Chan, D.C.; Chen, Y.L. Potential Therapeutic role of hispidulin in gastric cancer through induction of apoptosis via NAG-1 signaling. Evid. Based Complement. Altern. Med. 2013, 2013, 518301. [Google Scholar] [CrossRef]
  213. Anticancer Agents from Natural Products; Cragg, G.; Kingston, D.G.I.; Newman, D.J. (Eds.) CRC Press: Boca Raton, FL, USA; Taylor & Francis: Abingdon, UK, 2012. [Google Scholar]
  214. Newman, D.J.; Cragg, G.M. Natural Products as sources of new drugs from 1981 to 2014. J. Nat. Prod. 2016, 79, 629–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Mishra, B.B.; Tiwari, V.K. Natural products: An evolving role in future drug discovery. Eur. J. Med. Chem. 2011, 46, 4769–4807. [Google Scholar] [CrossRef]
  216. Safe, S.; Abbruzzese, J.; Abdelrahim, M.; Hedrick, E. Specificity protein transcription factors and cancer: Opportunities for drug development. Cancer Prev. Res. Phila 2018, 11, 371–382. [Google Scholar] [CrossRef] [PubMed]
  217. Lee, S.H.; Cekanova, M.; Baek, S.J. Multiple mechanisms are involved in 6-gingerol-induced cell growth arrest and apoptosis in human colorectal cancer cells. Mol. Carcinog. 2008, 47, 197–208. [Google Scholar] [CrossRef] [Green Version]
  218. Piyanuch, R.; Sukhthankar, M.; Wandee, G.; Baek, S.J. Berberine, a natural isoquinoline alkaloid, induces NAG-1 and ATF3 expression in human colorectal cancer cells. Cancer Lett. 2007, 258, 230–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Auyeung, K.K.; Ko, J.K. Coptis chinensis inhibits hepatocellular carcinoma cell growth through nonsteroidal anti-inflammatory drug-activated gene activation. Int. J. Mol. Med. 2009, 24, 571–577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Nualsanit, T.; Rojanapanthu, P.; Gritsanapan, W.; Lee, S.H.; Lawson, D.; Baek, S.J. Damnacanthal, a noni component, exhibits antitumorigenic activity in human colorectal cancer cells. J. Nutr. Biochem. 2012, 23, 915–923. [Google Scholar] [CrossRef] [Green Version]
  221. Bottone, F.G., Jr.; Baek, S.J.; Nixon, J.B.; Eling, T.E. Diallyl disulfide (DADS) induces the antitumorigenic NSAID-activated gene (NAG-1) by a p53-dependent mechanism in human colorectal HCT 116 cells. J. Nutr. 2002, 132, 773–778. [Google Scholar] [CrossRef] [Green Version]
  222. Lee, S.-H.; Kim, J.-S.; Yamaguchi, K.; Eling, T.E.; Baek, S.J. Indole-3-carbinol and 3,3′-diindolylmethane induce expression of NAG-1 in a p53-independent manner. Biochem. Biophys. Res. Commun. 2005, 328, 63–69. [Google Scholar] [CrossRef]
  223. Baek, S.J.; Kim, J.-S.; Jackson, F.R.; Eling, T.E.; McEntee, M.F.; Lee, S.-H. Epicatechin gallate-induced expression of NAG-1 is associated with growth inhibition and apoptosis in colon cancer cells. Carcinogenesis 2004, 25, 2425–2432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Cho, K.-N.; Sukhthankar, M.; Lee, S.-H.; Yoon, J.-H.; Baek, S.J. Green tea catechin (−)-epicatechin gallate induces tumour suppressor protein ATF3 via EGR-1 activation. Eur. J. Cancer 2007, 43, 2404–2412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Wilson, L.C.; Baek, S.J.; Call, A.; Eling, T.E. Nonsteroidal anti-inflammatory drug-activated gene (NAG-1) is induced by genistein through the expression of p53 in colorectal cancer cells. Int. J. Cancer 2003, 105, 747–753. [Google Scholar] [CrossRef]
  226. Kim, C.H.; Kim, M.Y.; Moon, J.Y.; Hwang, J.W.; Lee, S.Y.; Joo, Y.M.; Han, S.I.; Park, H.G.; Kang, H.S. Implication of NAG-1 in synergistic induction of apoptosis by combined treatment of sodium salicylate and PI3K/MEK1/2 inhibitors in A549 human lung adenocarcinoma cells. Biochem. Pharm. 2008, 75, 1751–1760. [Google Scholar] [CrossRef] [PubMed]
  227. Tsai, S.F.; Tao, M.; Ho, L.I.; Chiou, T.W.; Lin, S.Z.; Su, H.L.; Harn, H.J. Isochaihulactone-induced DDIT3 causes ER stress-PERK independent apoptosis in glioblastoma multiforme cells. Oncotarget 2017, 8, 4051–4061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Chen, Y.L.; Lin, P.C.; Chen, S.P.; Lin, C.C.; Tsai, N.M.; Cheng, Y.L.; Chang, W.L.; Lin, S.Z.; Harn, H.J. Activation of nonsteroidal anti-inflammatory drug-activated gene-1 via extracellular signal-regulated kinase 1/2 mitogen-activated protein kinase revealed a isochaihulactone-triggered apoptotic pathway in human lung cancer A549 cells. J. Pharm. Exp. Ther. 2007, 323, 746–756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Shin, D.Y.; Kim, G.Y.; Li, W.; Choi, B.T.; Kim, N.D.; Kang, H.S.; Choi, Y.H. Implication of intracellular ROS formation, caspase-3 activation and Egr-1 induction in platycodon D-induced apoptosis of U937 human leukemia cells. Biomed. Pharm. 2009, 63, 86–94. [Google Scholar] [CrossRef]
  230. Ko, J.K.; Leung, W.C.; Ho, W.K.; Chiu, P. Herbal diterpenoids induce growth arrest and apoptosis in colon cancer cells with increased expression of the nonsteroidal anti-inflammatory drug-activated gene. Eur. J. Pharm. 2007, 559, 1–13. [Google Scholar] [CrossRef]
  231. Lim, J.H.; Park, J.W.; Min, D.S.; Chang, J.S.; Lee, Y.H.; Park, Y.B.; Choi, K.S.; Kwon, T.K. NAG-1 up-regulation mediated by EGR-1 and p53 is critical for quercetin-induced apoptosis in HCT116 colon carcinoma cells. Apoptosis 2007, 12, 411–421. [Google Scholar] [CrossRef]
  232. Baek, S.J.; Wilson, L.C.; Eling, T.E. Resveratrol enhances the expression of non-steroidal anti-inflammatory drug-activated gene (NAG-1) by increasing the expression of p53. Carcinogenesis 2002, 23, 425–432. [Google Scholar] [CrossRef] [Green Version]
  233. Golkar, L.; Ding, X.Z.; Ujiki, M.B.; Salabat, M.R.; Kelly, D.L.; Scholtens, D.; Fought, A.J.; Bentrem, D.J.; Talamonti, M.S.; Bell, R.H.; et al. Resveratrol inhibits pancreatic cancer cell proliferation through transcriptional induction of macrophage inhibitory cytokine-1. J. Surg. Res. 2007, 138, 163–169. [Google Scholar] [CrossRef]
  234. 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]
  235. Jutooru, I.; Chadalapaka, G.; Abdelrahim, M.; Basha, M.R.; Samudio, I.; Konopleva, M.; Andreeff, M.; Safe, S. Methyl 2-cyano-3,12-dioxooleana-1,9-dien-28-oate decreases specificity protein transcription factors and inhibits pancreatic tumor growth: Role of microRNA-27a. Mol. Pharmacol. 2010, 78, 226–236. [Google Scholar] [CrossRef] [Green Version]
  236. Jutooru, I.; Chadalapaka, G.; Lei, P.; Safe, S. Inhibition of NFkappaB and pancreatic cancer cell and tumor growth by curcumin is dependent on specificity protein down-regulation. J. Biol. Chem. 2010, 285, 25332–25344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Chintharlapalli, S.; Papineni, S.; Lei, P.; Pathi, S.; Safe, S. Betulinic acid inhibits colon cancer cell and tumor growth and induces proteasome-dependent and -independent downregulation of specificity proteins (Sp) transcription factors. BMC Cancer 2011, 11, 371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Gandhy, S.U.; Kim, K.; Larsen, L.; Rosengren, R.J.; Safe, S. Curcumin and synthetic analogs induce reactive oxygen species and decreases specificity protein (Sp) transcription factors by targeting microRNAs. BMC Cancer 2012, 12, 564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Jutooru, I.; Chadalapaka, G.; Sreevalsan, S.; Lei, P.; Barhoumi, R.; Burghardt, R.; Safe, S. Arsenic trioxide downregulates specificity protein (Sp) transcription factors and inhibits bladder cancer cell and tumor growth. Exp. Cell Res. 2010, 316, 2174–2188. [Google Scholar] [CrossRef] [Green Version]
  240. Pathi, S.S.; Lei, P.; Sreevalsan, S.; Chadalapaka, G.; Jutooru, I.; Safe, S. Pharmacologic doses of ascorbic acid repress specificity protein (Sp) transcription factors and Sp-regulated genes in colon cancer cells. Nutr. Cancer 2011, 63, 1133–1142. [Google Scholar] [CrossRef] [Green Version]
  241. Lou, Z.; O’Reilly, S.; Liang, H.; Maher, V.M.; Sleight, S.D.; McCormick, J.J. Down-regulation of overexpressed sp1 protein in human fibrosarcoma cell lines inhibits tumor formation. Cancer Res. 2005, 65, 1007–1017. [Google Scholar] [PubMed]
  242. Hedrick, E.; Cheng, Y.; Jin, U.H.; Kim, K.; Safe, S. Specificity protein (Sp) transcription factors Sp1, Sp3 and Sp4 are non-oncogene addiction genes in cancer cells. Oncotarget 2016, 7, 22245–22256. [Google Scholar] [CrossRef] [PubMed]
  243. Scott, G.K.; Mattie, M.D.; Berger, C.E.; Benz, S.C.; Benz, C.C. Rapid alteration of microRNA levels by histone deacetylase inhibition. Cancer Res. 2006, 66, 1277–1281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Tillotson, L.G. RIN ZF, a novel zinc finger gene, encodes proteins that bind to the CACC element of the gastrin promoter. J. Biol. Chem. 1999, 274, 8123–8128. [Google Scholar] [CrossRef] [Green Version]
  245. Mertens-Talcott, S.U.; Chintharlapalli, S.; Li, X.; Safe, S. The oncogenic microRNA-27a targets genes that regulate specificity protein transcription factors and the G2-M checkpoint in MDA-MB-231 breast cancer cells. Cancer Res. 2007, 67, 11001–11011. [Google Scholar] [CrossRef] [Green Version]
  246. Kim, K.; Chadalapaka, G.; Lee, S.O.; Yamada, D.; Sastre-Garau, X.; Defossez, P.A.; Park, Y.Y.; Lee, J.S.; Safe, S. Identification of oncogenic microRNA-17-92/ZBTB4/specificity protein axis in breast cancer. Oncogene 2012, 31, 1034–1044. [Google Scholar] [CrossRef] [Green Version]
  247. Chadalapaka, G.; Jutooru, I.; Safe, S. Celastrol decreases specificity proteins (Sp) and fibroblast growth factor receptor-3 (FGFR3) in bladder cancer cells. Carcinogenesis 2012, 33, 886–894. [Google Scholar] [CrossRef] [Green Version]
  248. Jutooru, I.; Guthrie, A.S.; Chadalapaka, G.; Pathi, S.; Kim, K.; Burghardt, R.; Jin, U.H.; Safe, S. Mechanism of action of phenethylisothiocyanate and other reactive oxygen species-inducing anticancer agents. Mol. Cell. Biol. 2014, 34, 2382–2395. [Google Scholar] [CrossRef] [Green Version]
  249. Kasiappan, R.; Jutooru, I.; Karki, K.; Hedrick, E.; Safe, S. Benzyl Isothiocyanate (BITC) induces reactive oxygen species-dependent repression of STAT3 Protein by down-regulation of specificity proteins in pancreatic cancer. J. Biol. Chem. 2016, 291, 27122–27133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Karki, K.; Hedrick, E.; Kasiappan, R.; Jin, U.H.; Safe, S. Piperlongumine induces Reactive Oxygen Species (ROS)-Dependent Downregulation of specificity protein transcription factors. Cancer Prev. Res. Phila 2017, 10, 467–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. O’Hagan, H.M.; Wang, W.; Sen, S.; Destefano Shields, C.; Lee, S.S.; Zhang, Y.W.; Clements, E.G.; Cai, Y.; Van Neste, L.; Easwaran, H.; et al. Oxidative damage targets complexes containing DNA methyltransferases, SIRT1, and polycomb members to promoter CpG Islands. Cancer Cell 2011, 20, 606–619. [Google Scholar] [CrossRef] [Green Version]
  252. Liang, Z.; Zhang, B.; Su, W.W.; Williams, P.G.; Li, Q.X. C-Glycosylflavones Alleviate Tau Phosphorylation and Amyloid Neurotoxicity through GSK3beta Inhibition. ACS Chem. Neurosci. 2016, 7, 912–923. [Google Scholar] [CrossRef] [PubMed]
  253. Liang, Z.; Li, Q.X. Discovery of selective, substrate-competitive, and passive membrane permeable glycogen synthase kinase-3beta inhibitors: Synthesis, biological evaluation, and molecular modeling of new C-Glycosylflavones. ACS Chem. Neurosci. 2018, 9, 1166–1183. [Google Scholar] [CrossRef] [PubMed]
  254. Tan, X.; Liang, Z.; Li, Y.; Zhi, Y.; Yi, L.; Bai, S.; Forest, K.H.; Nichols, R.A.; Dong, Y.; Li, Q.X. Isoorientin, a GSK-3β inhibitor, rescues synaptic dysfunction, spatial memory deficits and attenuates pathological progression in APP/PS1 model mice. Behav. Brain Res. 2021, 398, 112968. [Google Scholar] [CrossRef] [PubMed]
  255. Doello, S.; Liang, Z.; Cho, I.K.; Kim, J.B.; Li, Q.X. Cytotoxic effects of 24-Methylenecyloartanyl ferulate on A549 nonsmall cell lung cancer cells through MYBBP1A up-regulation and AKT and aurora B kinase inhibition. J. Agric. Food Chem. 2018, 66, 3726–3733. [Google Scholar] [CrossRef]
  256. Kim, H.; Kim, D.; Jang, G.; Lee, S.; Jang, H.; Cho, I.; Li, Q.X.; Lee, S.; Kim, J. Purification of cycloartenyl ferulate, 24-methylenecycloartanyl ferulate, campesteryl ferulate and sitosteryl ferulate from rice bran and their effects on the NLRP3 inflammasome. Acad. J. Agric. Res. 2016, 4, 411–419. [Google Scholar] [CrossRef]
  257. Chang, C.L.; Cho, I.K.; Li, Q.X. Insecticidal activity of basil oil, trans-anethole, estragole, and linalool to adult fruit flies of Ceratitis capitata, Bactrocera dorsalis, and Bactrocera cucurbitae. J. Econ. Entomol. 2009, 102, 203–209. [Google Scholar] [CrossRef]
  258. Li, A.S.; Iijima, A.; Huang, J.; Li, Q.X.; Chen, Y. Putative mode of action of the monoterpenoids linalool, methyl eugenol, estragole, and citronellal on ligand-gated ion channels. Engineering 2020, 6, 541–545. [Google Scholar] [CrossRef]
  259. Sun, B.T.D.; Pan, D.; Baker, M.R.; Liang, Z.; Wang, Z.; Lei, J.; Liu, S.; Hu, C.Y.; Li, Q.X. Dihydromyricetin imbues anti-adipogenic effects on 3T3-L1 cells via direct interactions with 78-kDa glucose regulated protein. J. Nutr. 2021. [Google Scholar] [CrossRef]
  260. Boon, C.S.; Mcclements, D.J.; Weiss, J.; Decker, E.A. Factors influencing the chemical stability of carotenoids in foods. Crit. Rev. Food Sci. Nutr. 2010, 50, 515–532. [Google Scholar] [CrossRef] [PubMed]
  261. Pan, K.; Zhong, Q. Organic nanoparticles in foods: Fabrication, characterization, and utilization. Annu. Rev. Food Sci. Technol. 2016, 7, 245–266. [Google Scholar] [CrossRef] [PubMed]
  262. Nile, S.H.; Baskar, V.; Selvaraj, D.; Nile, A.; Xiao, J.; Kai, G. Nanotechnologies in food science: Applications, recent trends, and future perspectives. Nano-Micro Lett. 2020, 12, 45. [Google Scholar] [CrossRef] [Green Version]
  263. McClements, D.J. Nanotechnology approaches for improving the healthiness and sustainability of the modern food supply. ACS Omega 2020, 5, 29623–29630. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Cyclooxygenase-derived prostaglandins regulate T helper cell differentiation and function. Dendritic cells (DCs) present antigens to naïve CD4+ T cells and produce cytokines which induce T helper cell differentiation to Th1, Th2, Th9 and Th17 cell subsets. COX-derived eicosanoids can affect Th differentiation and function in multiple ways. Thromboxane A2 (TXA2) inhibits the DC/T cell interactions and reduces T cell differentiation. PGE2 and PGI2 inhibit production of IL-12 which results in reduced Th1 differentiation and indirectly promotes Th2 differentiation. PGD2 and PGE2 promote Th9 differentiation while PGF and PGI2 suppress Th17 differentiation. Together, these eicosanoids regulate the immune response.
Figure 1. Cyclooxygenase-derived prostaglandins regulate T helper cell differentiation and function. Dendritic cells (DCs) present antigens to naïve CD4+ T cells and produce cytokines which induce T helper cell differentiation to Th1, Th2, Th9 and Th17 cell subsets. COX-derived eicosanoids can affect Th differentiation and function in multiple ways. Thromboxane A2 (TXA2) inhibits the DC/T cell interactions and reduces T cell differentiation. PGE2 and PGI2 inhibit production of IL-12 which results in reduced Th1 differentiation and indirectly promotes Th2 differentiation. PGD2 and PGE2 promote Th9 differentiation while PGF and PGI2 suppress Th17 differentiation. Together, these eicosanoids regulate the immune response.
Nutrients 13 01881 g001
Figure 2. Mechanism of ROS-inducing compounds (ROS-Ic) in cancer cells; includes induction of ROS, decreased expression of Myc and Myc-regulated miRs, induction of ZBTBs and downregulation of Sp transcription factors and Sp-regulated genes (216).
Figure 2. Mechanism of ROS-inducing compounds (ROS-Ic) in cancer cells; includes induction of ROS, decreased expression of Myc and Myc-regulated miRs, induction of ZBTBs and downregulation of Sp transcription factors and Sp-regulated genes (216).
Nutrients 13 01881 g002
Table 1. Phytochemicals that modulate the integrity of the blood–brain barrier after ischemia.
Table 1. Phytochemicals that modulate the integrity of the blood–brain barrier after ischemia.
PhytochemicalAnimalMechanism of ActionReference
Ascorbic acidRatDownregulation of MMP-2 and MMP-9[145]
Astragaloside IVRatDownregulation of MMP-9 and AQP4[147]
BaicalinRatDownregulation of MMP-9[152]
Chlorogenic acidRatDownregulation of MMP-2 and MMP-9[140]
CrocinRatDownregulation of MMP-2 and MMP-9[141]
CurcuminRatDownregulation of MMP-9[153]
Dl-3-n-butylphthalideMouseDownregulation of Caveolin-1[154]
Ellagic acidRatDownregulation of AQP4 and MMP-9[148]
GastrodinRatDownregulation of MMP-2 and MMP-9[144]
Ginsenoside Rb1MouseDownregulation of MMP-9[155]
Ginsenoside RdRatDownregulation of NF-κB and MMP-9[156]
HesperidinMouseInhibition of FoxO3a nuclear translocation
Downregulation of MMP-3/9
[150]
Icariside IIRatDownregulation of MMP-9
Upregulation of TIMP-1
[157]
JuglaninMouseDownregulation of VEGF and VEGFR2[151]
MelatoninRatDownregulation of MMP-9[158]
PinocembrinRatDownregulation of MMP-2 and MMP-9[142]
QuercetinRatDownregulation of MMP-9[159]
ResveratrolRatDownregulation of MMP-9
Upregulation of TIMP-1
[160]
RutinRatDownregulation of MMP-9[161]
Salvianolic acid ARatDownregulation ofMMP-9
Upregulation of TIMP-1
Src phosphorylation at Tyr416
[143,146]
Sodium tanshinone IIA sulfonate (with rt-PA)HumanDownregulation of MMP-9 and TIMP-1[139]
TetrahydrocurcuminMouseDownregulation of MMP-9
Upregulation of TIMP-2
[162]
TetramethylpyrazineRatDownregulation of JAK/STAT phosphorylation[163]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Baek, S.J.; Hammock, B.D.; Hwang, I.-K.; Li, Q.X.; Moustaid-Moussa, N.; Park, Y.; Safe, S.; Suh, N.; Yi, S.-S.; Zeldin, D.C.; et al. Natural Products in the Prevention of Metabolic Diseases: Lessons Learned from the 20th KAST Frontier Scientists Workshop. Nutrients 2021, 13, 1881. https://doi.org/10.3390/nu13061881

AMA Style

Baek SJ, Hammock BD, Hwang I-K, Li QX, Moustaid-Moussa N, Park Y, Safe S, Suh N, Yi S-S, Zeldin DC, et al. Natural Products in the Prevention of Metabolic Diseases: Lessons Learned from the 20th KAST Frontier Scientists Workshop. Nutrients. 2021; 13(6):1881. https://doi.org/10.3390/nu13061881

Chicago/Turabian Style

Baek, Seung J., Bruce D. Hammock, In-Koo Hwang, Qing X. Li, Naima Moustaid-Moussa, Yeonhwa Park, Stephen Safe, Nanjoo Suh, Sun-Shin Yi, Darryl C. Zeldin, and et al. 2021. "Natural Products in the Prevention of Metabolic Diseases: Lessons Learned from the 20th KAST Frontier Scientists Workshop" Nutrients 13, no. 6: 1881. https://doi.org/10.3390/nu13061881

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